You are on page 1of 10

Kinematic Framework for Evaluating Seismic Earth

Pressures on Retaining Walls


Scott J. Brandenberg, M.ASCE 1; George Mylonakis, M.ASCE 2; and Jonathan P. Stewart, F.ASCE 3

Abstract: During earthquake ground shaking earth pressures on retaining structures can cyclically increase and decrease as a result of inertial
forces applied to the walls and kinematic interactions between the stiff wall elements and surrounding soil. The application, based on limit
equilibrium analysis, of a pseudostatic inertial force to a soil wedge behind the wall [the mechanism behind the widely-used Mononobe–
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

Okabe (M–O) method] is a poor analogy for either inertial or kinematic wall–soil interaction. This paper demonstrates that the kinematic
component of interaction varies strongly with the ratio of wavelength to wall height (λ=H), asymptotically approaching zero for large λ=H,
and oscillating between the peak value and zero for λ=H < 2.3. Base compliance, represented in the form of translational and rotational
stiffness, reduces seismic earth pressure by permitting the walls to conform more closely to the free-field soil displacement profile. This
framework can explain both relatively low seismic pressures relative to M–O predictions observed in recent experiments with λ=H > ∼10,
and relatively high seismic earth pressures relative to M–O from numerical analyses in the literature with λ=H ¼ 4. DOI: 10.1061/(ASCE)
GT.1943-5606.0001312. © 2015 American Society of Civil Engineers.
Author keywords: Wall; Seismic earth pressure; Wave; Analysis; Dynamic testing.

Introduction Recent work based on experiments and various dynamic solu-


tions considering elastic soil behavior has, directly or indirectly,
The increment of lateral earth pressure that should be applied dur- challenged this practice as being both too conservative (e.g., Al
ing the design of retaining walls to account for earthquake effects Atik and Sitar 2010; Lew et al. 2010) and as being un-conservative
has been a source of confusion among design professionals and a (e.g., Wood 1973; Veletsos and Younan 1994; Ostadan 2005).
topic on which there are divergent opinions among researchers. These conflicting findings, based on different approaches and as-
Current guidelines documents (e.g., NCHRP 2008) prescribe sub- sumptions regarding system behavior, drive a good deal of the
stantial seismic earth pressures beyond those for the preseismic confusion on the subject of seismic earth pressures on retaining
(generally active) condition. These recommendations are based on walls. A fundamental problem is that the M–O method does not
a limit equilibrium analysis in which a pseudostatic seismic coef- adequately represent interaction of vibrating soil in the free field
ficient (kh ) acts upon an active Coulomb-type wedge in frictional with an embedded structure or a retaining wall. This interaction
soil, which in turn results in an incremental change in the lateral may be best understood using a conceptual framework, rooted in
force applied to the wall, PAE , over its static counterpart PA . This the principles of soil–structure interaction and wave propagation, in
approach is based on the classical work by Okabe (1924) and which kinematic and inertial interaction effects are distinguished.
Mononobe and Matsuo (1929), widely known as the Mononobe– The next section describes a conceptual framework for defining
Okabe (M–O) method, with modest modification by Seed and seismic earth pressures from kinematic interaction in terms of the
Whitman (1970). More accurate variants on the classical approach ratio of wavelength of vertically propagating shear waves to wall
using nonplanar failure surfaces (Chen 1975; Chen and Liu 1990) height. This approach convincingly explains the apparently diver-
and approximate accounting for the phasing of inertial demands gent findings from centrifuge tests by Al Atik and Sitar (2010) and
within the wedge (Steedman and Zeng 1990) are conceptually alike the numerical results from Ostadan (2005). Recommendations for
and provide similar results for the active case. rational simplified analysis of seismic earth pressures in engineer-
ing practice are then presented, along with conditions for which
1 more elaborate analyses are needed.
Associate Professor and Vice Chair, Dept. of Civil and Environmental
Engineering, 5731 Boelter Hall, Univ. of California, Los Angeles,
CA 90095-1593 (corresponding author). E-mail: sjbrandenberg@ucla.edu
2
Professor and Chair in Geotechnics and Soil–Structure Interaction, Conceptual Framework
Dept. of Civil Engineering, University Walk, Clifton BS8, Univ. of Bristol,
U.K.; Professor, Univ. of Patras, Greece; Adjunct Professor, Dept. of Civil The seismic increment to lateral earth pressures can be considered
and Environmental Engineering, 5731 Boelter Hall, Univ. of California, as having kinematic and inertial components, as illustrated in Fig. 1
Los Angeles, CA 90095-1593. E-mail: g.mylonakis@bristol.ac.uk for an embedded building foundation with relatively stiff basement
3
Professor and Chair, Dept. of Civil and Environmental Engineering, walls. The free-field motion imposed on this system (ug ) varies
5731 Boelter Hall, Univ. of California, Los Angeles, CA 90095-1593. with depth as indicated in Fig. 1(a). In the kinematic problem
E-mail: jstewart@seas.ucla.edu
for which there is no structure or wall inertia, the motion of the
Note. This manuscript was submitted on July 18, 2013; approved on
January 30, 2015; published online on March 16, 2015. Discussion period
foundation at base depth H is denoted uFIM [in reference to foun-
open until August 16, 2015; separate discussions must be submitted for dation input motion (FIM)], which differs from the free-field mo-
individual papers. This paper is part of the Journal of Geotechnical tion at this same depth, ug ðHÞ, as a result of relative foundation/
and Geoenvironmental Engineering, © ASCE, ISSN 1090-0241/ free-field displacements associated with wall–soil contact stresses,
04015031(10)/$25.00. as well as base slab averaging effects that occur in the presence of

© ASCE 04015031-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


Fig. 1. Schematic illustration of the following: (a) kinematic interaction components of foundation–soil interaction for an embedded foundation
system; (b) application of free-field motion, modified for base slab averaging effects ugm , for inertial interaction
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

inclined or incoherent waves (e.g., Veletsos and Prasad 1989). The is a simplifying assumption commonly used in soil–structure inter-
kinematic component of seismic earth pressures accounts for the action problems because it permits development of tractable solu-
interaction between the free-field motion ug ðzÞ and the structural tions. The values assigned to the stiffness terms should account for
wall elements, apart from their inertia and any external inertial coupling between various foundation vibrations modes, as de-
loads imposed upon the system. scribed in a subsequent paragraph. Although these assumptions
As shown in Fig. 1(b), the inertial interaction problem involves may appear to be limiting, the method can be readily extended
computation of the response of a structure and its foundation to to a wide range of practical conditions (including nonrigid founda-
the kinematic ground motions. Inertial forces from the structure tions as well as nonlinear and nonuniform soil) in a manner typical
cause additional relative displacements between the foundation of soil–structure interaction applications (NIST 2012) as illustrated
and the free-field, and additional increments of seismic earth pres- subsequently.
sure. The springs and dashpots in Fig. 1(b) represent the impedance The model derivation is described in two stages, as follows:
of the foundation from translation and rocking vibration modes (1) wall pressures and their resultant demands (forces and mo-
(e.g., Pais and Kausel 1988; Gazetas 1991). ments) are derived from the product of differential wall/free-field
In light of the previously mentioned soil–structure interaction displacement and wall-soil stiffness, and (2) equations for the stiff-
framework, the soil wedge concept currently used to evaluate seis- ness terms are developed (which is essential for analysis of force/
mic earth pressures will seldom have relevance to the physical moment demands and differential wall/free-field motions). Funda-
mechanisms producing those pressures. Even in cases where a state mental characteristics of wall–soil interaction derived from these
of active earth pressure (and its associated soil wedge) exists prior analyses are then described and illustrated using example solutions,
to seismic shaking, increments of earth pressure from earthquake which demonstrate that the wall–soil interaction response depends
ground shaking will arise from relative displacements between the strongly on the ratio of wavelength to wall height.
wall and free-field soil associated with kinematic and inertial inter-
action, which is not well-represented by a seismic coefficient acting
on an active wedge. Inertial interaction can mobilize large relative Wall–Soil Interaction Forces and Displacements
displacements when, for example, a massive structure is connected
to the wall elements and base shear mobilizes reaction stresses at A rigid U-shaped structure with vertical walls embedded in a soil
the soil/wall interface. Such effects can be evaluated as part of seis- profile experiencing vertically propagating harmonic free-field
mic structural response analysis if soil springs are included in the shear waves is shown in Fig. 2. The free-field ground motion is
structural model. Free-standing walls or basement walls not struc-
turally connected to lateral force resisting elements in structures
would have seismic earth pressures dominated by kinematic inter-
action, which is the topic addressed in the remainder of this paper.

Model Derivation

Seismic earth pressures arising from kinematic interaction are for-


mulated based on the following assumptions (Fig. 2): (1) infinitely
long U-shaped structure with rigid walls and rigid base slab is em-
bedded in a soil profile with a uniform shear wave velocity; (2) a
horizontally coherent vertically propagating shear wave interacts
with the embedded structure and base slab averaging effects are
negligible; (3) the soil and wall are in perfect contact, and a gap
does not form at this interface; and (4) the interaction between the
Fig. 2. Schematic of embedded rigid strip foundation excited by ver-
soil and vertical walls is characterized by stiffness intensity terms,
tically propagating shear wave. This schematic strictly holds for a case
kiy and kiz (defined in the subsequent text), and interaction between
of no base slab averaging. Such effects can be considered by multiply-
the soil and base slab is characterized by stiffness terms K y and
ing ug ðzÞ by an appropriate base slab averaging transfer function to
K xx . These stiffness terms satisfy the Winkler assumption that
obtain ugm ðzÞ (e.g., NIST 2012)
the stiffness values act independently from one another, which

© ASCE 04015031-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


consistent with the influence of the free-surface since the shear where uFIM and θFIM = base slab translation and rotation,
strain is zero at z ¼ 0. respectively.
Kinematic wall pressures arise from incompatibility in the For a rigid wall resting on a rigid base, θFIM must be zero, base
displacement of the rigid wall and the free-field soil column. displacement must equal free-field displacement at the base of the
Accordingly, the integral of the horizontal stress increment over wall (i.e., uFIM ¼ ug0 cos kH), and the solution for PE and M E may
the height of the wall is the kinematic seismic force increment easily be obtained from Eqs. (1) and (2) for a free-field ground
PE (PE is adopted in this paper instead of PAE , which is associated motion with any particular wavelength. However, a more general
with M–O theory, because the solution developed in this paper solution for a wall embedded within an elastic layer, thereby ex-
does not require an active condition). For ground motion in the hibiting base compliance, can also be obtained. The rotational stiff-
y-direction, PE is calculated as a force per unit length ness of the embedded strip foundation contains contributions from
Z H the base slab and from vertical shear tractions and horizontal nor-
PE ¼ kiy ½ug0 cos kz − uw ðzÞdz ð1Þ mal stresses acting on the walls. The horizontal stresses acting on
0 the walls are explicitly included in Eqs. (1) and (2).
To solve for the foundation input motions, horizontal force and
where H = wall height; uw ðzÞ = wall displacement at depth z; kiy =
moment equilibrium of the foundation slab are considered, assum-
soil–wall reaction stiffness in y-direction (normal stresses) per unit
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

ing that the free-field ground motion is input to the free-ends of the
of wall area (superscript i denotes stiffness intensity measured in
spring elements. Substituting Eq. (4) into Eqs. (1) and (2), and re-
units of F=L3 ; details in the subsequent text); k ¼ 2π=λ = wave
quiring horizontal force and moment equilibrium between the wall
number; and λ = wavelength of the shear wave propagating
resultants and base reactions provides
vertically through the soil. The moment applied by the horizontal
Z H
soil–wall interaction stresses relative to the foundation slab base
elevation is PE ¼ kiy ½ug0 cos kz − uFIM − θFIM ðH − zÞdz
0
Z H Ky
ME ¼ kiy ðH − zÞ½ug0 cos kz − uw ðzÞdz ð2Þ ¼ ðu − ug0 cos kHÞ ð5aÞ
0
2 FIM
Z H
Eqs. (1) and (2) can be combined to calculate the location of ME ¼ kiy ½ug0 cos kz − uFIM − θFIM ðH − zÞðH − zÞdz
resultant PE , measured as distance h upwards from the base of 0
the wall as K xx θFIM
¼ þ kiz HB2 θFIM ð5bÞ
2
h M
¼ E ð3Þ Stiffness terms K y and K xx are multiplied by one-half to account
H PE H
for two vertical walls being attached to a single rigid base. The kiz
The depth-dependent wall displacement uw ðzÞ for a rigid wall term on the right side of Eq. (5b) represents the moment induced by
and foundation system is vertical tractions acting on the walls. By evaluating the integrals
and rearranging terms, the subsequent solution is obtained for
uw ðzÞ ¼ uFIM þ θFIM ðH − zÞ ð4Þ foundation displacements

uFIM ½6H2 ðkiy Þ2 þ 3k2 K y ðK xx þ 2kiz HB2 Þ þ 2k2 H3 K y kiy  cos kH þ ½4kH 3 ðkiy Þ2 þ 6kðK xx þ 2kiz HB2 Þkiy  sin kH − 6H2 ðkiy Þ2
Hu ¼ ¼
ug0 k2 ½H4 ðkiy Þ2 þ 2H3 K y kiy þ 6Hkiy þ 3K y ðK xx þ 2kiz HB2 Þ
ð6aÞ

θFIM B ½6k2 H2 K y kiy þ 24Hðkiy Þ2 þ 12K y kiy sin2 ðkH=2Þ − 6kH 2 ðkiy Þ2 sin kH − 3H2 k2 K y kiy
Hθ ¼ ¼B ð6bÞ
ug0 k2 ½H4 ðkiy Þ2 þ 2H3 K y kiy þ 6Hkiy þ 3K y ðK xx þ 2kiz HB2 Þ

These foundation displacements can then be inserted into and the base slab. Such partitioning is required to obtain the distri-
Eqs. (5a) and (5b) to obtain PE and M E for a compliant base con- bution of earth pressure acting on the vertical walls, which is our
dition. Varun et al. (2009) developed a solution for the kinematic objective. To overcome this problem, available solutions are first
displacements of axisymmetric caissons that is analogous to the used to define stiffness terms for individual foundation components
wall displacements derived here. under the assumption of no interaction between vibration modes
(i.e., the components are independent). Next, modification factors
χy and χxx are introduced to account for interaction between the
Stiffness of Wall–Soil System translation and rotation terms, respectively, such that the resulting
Having formulated the solution for PE and ME , the stiffness terms, global foundation stiffness matches published equations for em-
kiy , kiz , K y , and K xx are now evaluated. Classical inertial soil- bedded foundations. For simplicity, the base and wall stiffnesses
structure interaction (SSI) literature (e.g., summarized by Gazetas are both modified by the same χy and χxx terms.
1983; Mylonakis et al. 2006; NIST 2012) provides equations for
the overall stiffness of embedded foundations representing the inter- Horizontal Wall–Soil Stiffness Intensity kiy
action of the soil with the entire foundation system, but the global Kloukinas et al. (2012) developed a simple analytical expression
stiffness is not partitioned into contributions from the vertical walls for kiy for kinematic interaction between rigid vertical walls and

© ASCE 04015031-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


an elastic soil layer resting atop a rigid base. After correction of The solution in Eq. (9a) does not extrapolate properly to an in-
their published expression (a clerical error involving omission of finitely thick elastic layer, for which the stiffness of a strip footing
the square root in the denominator) and including the multiplier, is zero. On the other hand, under such a condition the solution in
χy , the stiffness intensity is obtained as Eq. (9b) is exact (Mushkelishvili 1963).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
π G 2ωH 2 Derivation of Interaction Terms χ y and χ xx
ky ¼ χy pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
i
1− ð7Þ The previously mentioned component stiffnesses can be combined
ð1 − νÞð2 − νÞ H πV s to compute overall static stiffnesses for the embedded wall–soil
system in translation and rocking. For translation, the stiffness is
where ω = angular frequency (rad=s). Material damping can be in- 2kiy H þ K y , which includes contributions from the vertical walls
corporated into the solution by using complex shear modulus, and the base slab. For rotation, the stiffness is kiy H2 þ K xx þ
Gð1 þ i2ξÞ, and complex shear wave velocity, V s ð1 þ iξÞ, where 2kiz HB2 , which includes contributions from horizontal and vertical
ξ = percent material damping. Kloukinas et al. (2012) develop kin- earth pressures acting on the vertical walls and the rotational stiff-
ematic earth pressures for a rigid wall resting atop a rigid base, ness of the base slab.
whereas the solution developed in this paper corresponds to soil Values of χy and χxx were selected such that the global stiffness
profiles that are deeper and compliant under the wall, which is
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

of the foundation matches the equations for embedded (denoted by


applicable to more realistic conditions. Coupling term χy is evalu- the subscript emb) strip footings by Jakub and Roesset (1977)
ated subsequently.
For an ideally undamped medium, the square root on the right-    
2.1G B 1H 4H
hand side of Eq. (7) can be interpreted as a dynamic stiffness modi- Ky emb ¼ 1þ2 1þ 1þ ¼ 2kiy H þ K y
fier (often denoted by α) that accounts for frequency dependence 2−ν D 3B 3D
from soil inertia, with the corresponding dashpot equal to zero. At ð10aÞ
ω ¼ πV s =2H the dynamic modifier becomes zero and at higher
frequencies kiy becomes imaginary meaning that the spring acts    
πGB2 1B H 2H
as a dashpot. This phenomenon is directly related to the rigid base K xx emb ¼ 1þ 1þ 1þ
condition used in the solution, which only allows radiation damp- 2ð1 − νÞ 5D B 3D
ing (from wave propagation away from the foundation) beyond the ¼ kiy H2 þ K xx þ 2kiz HB2 ð10bÞ
so-called cutoff frequency (e.g., Elsabee and Morray 1977). For
realistic systems involving a compliant base condition, the cutoff Expressions for χy and χxx can be obtained by substituting
frequency transition is smoother, allowing waves to exist at a wider Eqs. (7), (8), (9a), and (9b) into Eqs. (10a) and (10b). Fig. 3
range of frequencies (Li 1999), and material damping results in presents the values of χy and χxx versus H=B for various values
nonzero real and imaginary components at all frequencies. Elsabee of D=H. The solutions by Jakub and Roesset are intended for
and Morray (1977) suggest simple expressions for handling these conditions in which D=B > 2 and H=B < 2=3, and may provide
problems for embedded circular foundations, but there is presently erroneous results for conditions outside these bounds. Extrapola-
no simple solution analogous to Eq. (7) to account for these effects tion is bounded by the Kloukinas et al. (2012) solution for
for two-dimensional (2D) vertical walls. D=H ¼ 1, in which case χy ¼ 1.0, and the half-space solution
when D=H → ∞, in which case χy ¼ 0.0. These bounds are pre-
Vertical Wall–Soil Stiffness Intensity kiz sented in Fig. 3, and interpolation from Fig. 3 is recommended for
In accordance with the method of Kloukinas et al. (2012), the Sup- D=H < 2 and D=H > 20 rather than the values of χy and χxx im-
plemental Data presents the derivation of an expression for stiffness plied by Eqs. (7), (8), (9a), (9b), (10a), and (10b).
intensity associated with vertical tractions acting on walls (soil-wall
reaction stiffness in z-direction from shear), kiz . The resulting ex-
pression is Eq. 8, which contains a multiplier, χxx , that modifies Characteristics of Wall–Soil Interaction Response
the vertical stiffness to account for interaction associated with base
rotation and translation and is evaluated in a later section. Fig. 4 shows solutions for PE computed using Eq. (5a) with the
rffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2ffi expression for foundation input motion given in Eqs. (6a) and
π 2 − ν G 2ωH (6b). Results are plotted for various values of K y =ðkiy HÞ and
kiz ¼ χxx 1− ð8Þ ðK xx Þ=ðkiy H2 =3Þ, representing approximately the relative contribu-
2 1−νH πV s
tions of the base slab and horizontal normal stresses acting on the
walls to horizontal and rotational stiffness, respectively. In addition
Base Slab Stiffness Terms K y and K xx to the cases with a compliant base, a rigid base case (K y and
Gazetas and Roesset (1976) developed simple analytical expres- K xx → ∞) is included for comparison. For a given λ=H, the high-
sions for the translational and rotational stiffness (K y and K xx , est values of PE occur for the rigid base case. The value of PE de-
respectively) of a rigid strip footing resting on the surface of a creases as K y and K xx decrease because a more flexible base
homogeneous elastic layer of finite thickness overlying a rigid base. condition results in less relative displacement between the wall
Applying the coupling terms χy and χxx and adjusting the soil and free-field soil along the wall height.
thickness term to be equal to the distance from the base slab to the The most important interval of λ=H in Fig. 4 for application to
rigid base (i.e., using D − H), results in typical structural configurations and earthquake ground motions is
  the portion to the right of the longest wavelength (lowest fre-
2.1G B quency) peak in PE , which occurs at λ=H ≈ 2.3. The importance
K y ¼ χy 1þ2 ð9aÞ
2−ν D−H of this interval stems from its likely proximity to energetic portions
of the ground motion spectrum, which occur at the site resonant
  frequency or at frequencies controlled by the seismic source and
πGB2 1 B
K xx ¼ χxx 1þ ð9bÞ path (which are typically higher than the site frequency for sites
2ð1 − νÞ 5D − H in sedimentary basins).

© ASCE 04015031-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


To support the assertion that the important portion of Fig. 4
is typically λ=H ≥ 2.3, consider first the case of free field seismic
energy that is dominated by site resonance. The site resonant fre-
quency corresponds to λ=D ¼ 4, which can be manipulated to
λ=H ¼ 4D=H. Since the thickness of the soil column generally
significantly exceeds the wall height (i.e., generally D » H), λ=H
will typically exceed 4, which falls well to the right of the lowest
frequency peak at λ=H ≥ 2.3. For this resonant condition, the
largest kinematic pressures occur when D ¼ H (i.e., base slab is
founded on stiff rock overlain by soil).
Free-field ground motions are often not dominated by a
fundamental-mode site response, particularly in sedimentary basins
where seismic velocities gradually increase with depth without
having a distinct impedance contrast. In such cases, the controlling
ground motion period can be estimated as the mean period (T m =
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

period at the centroid of the Fourier amplitude spectrum), which is


typically in the range of 0.3–0.5 s for earthquakes in active crustal
regions in the magnitude range of engineering interest (Rathje et al.
2004). The corresponding wavelenths (computed as λ ¼ V s T m )
will seldom place the applicable value of λ=H below the peak at
∼2.3 for typical values of wall height H.
Based on the previously mentioned considerations, the most
useful insights into kinematic wall pressures are gained by studying
the portion of the results in Fig. 4 for λ=H > ∼2.3. Kinematic pres-
Fig. 3. Translational and rotational static stiffness interaction factors: sures are clearly high near the peak at 2.3 due to large relative de-
(a) χy versus H=B; (b) χxx versus H=B formations of wall and soil. As λ=H increases beyond 2.3, PE
decreases rapidly. In the limiting case where λ=H → ∞, the de-
formed shape of the free-field soil profile would become vertical
and would precisely conform to the shape of the rigid wall, thereby
resulting in zero kinematic interaction. The peaks and troughs in PE
observed for λ=H < 2.3 are caused by alternation of the direction
of the horizontal stress increment acting along the wall height as
frequency changes.
Fig. 5 shows kinematic transfer functions H u and Hθ associated
with the solution for the foundation input motion of Eqs. (6a)
and (6b). The transfer functions are compared to the recommen-
dation by Kausel et al. (1978), who used an embedded cylinder
geometry, assumed uFIM ¼ ug ðHÞ (this is the same as assuming
K y → ∞) and approximated high frequency interaction (i.e., at low
λ=H) as constant with respect to frequency for simplicity. At large
λ=H, the H u values for the rigid base case agree perfectly with
Kausel et al. (1978), whereas base compliance results in increased
translation and rotation. The assumption that uFIM ¼ ug ðHÞ is
approximate, even in the presence of vertically propagating coher-
ent waves, due to the wall–soil interaction force PE that must be
balanced by deflection of the base slab spring. As H=B increases,
translation amplitude decreases and rotational amplitude increases
for a particular λ=H.

Recommended Methods of Implementation

The solution for PE in Eq. (5a) is a function of wave number, k, and


is therefore a function of frequency. The dependence of PE on
frequency can be captured with two methods, as follows: (1) a
frequency-domain (FD) solution that takes as input a time-series
of free-field ground surface displacement [ug0 ðtÞ] or free-field dis-
placement modified for base slab averaging effects [ugm ðtÞ] (only
the ug0 notation is retained subsequently for brevity, although either
motion is suitable for application), or (2) a single-frequency (SF)
Fig. 4. Normalized PE versus normalized wavelength λ=H for
solution that takes as input a particular free-field displacement (ug0 )
various contributions of wall normal stress to translational and ro-
and a single frequency anticipated to dominate dynamic earth pres-
tational stiffness represented as ðK xx þ 2kiz HB2 Þ=ðkiy H2 =3Þ ¼(a) 3;
sure response. Both methods will be useful in design applications
(b) 10; (c) 100
and are described next.

© ASCE 04015031-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


5. Compute the time series of the seismic earth pressure resultant
PE ðtÞ using the inverse fast Fourier-transform algorithm. Find
the maximum value of this time series. The total demand on
the wall is the sum of PE [at the location indicated by Eq. (3)]
and the resultant of the initial earth pressure (typically at
z ¼ 2H=3).
Each of the frequency-domain displacements and forces given
above is complex-valued.
The single-frequency solution (SF solution) follows a six-step
process:
1. Estimate the mean period (T m ) of the design earthquake
ground motion. For projects where ground motions are esti-
mated using site-specific probabilistic seismic hazard analysis
followed by the selection of applicable accelerograms, the
mean period can be computed for each record using proce-
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

dures given in Rathje et al. (2004). When such accelerograms


are unavailable, T m can be computed from applicable ground
motion prediction equations (e.g., Rathje et al. 2004), or in
cases of sites having significant impedance contrasts giving
rise to strongly resonant responses, from the site period
(T ¼ 4H=V s ).
2. Compute kiy , kiz , K y , K xx , K y emb , and K xx emb using Eqs. (7),
(8), (9a), (9b), (10a), and (10b), or alternate solutions as de-
scribed in Step 2 of the previous procedure. For many practical
Fig. 5. Kinematic transfer functions for the following: (a) translational
situations, static stiffnesses will suffice for these quantities
foundation input motions derived from the research reported in this (zero frequency), although more precision is possible through
paper and compared to the simplified approach of Kausel et al. (1978); consideration of frequency dependence.
(b) rotational foundation input motions derived from the research 3. Use the results in Fig. 4, or a site-specific solution of
reported in this paper and compared to the simplified approach of Eqs. (5a) and (5b), to evaluate the variation of normalized
Kausel et al. (1978) PE [i.e., jPE j=ðug0 kiy HÞ] versus λ=H.
4. Compute λ=H, based on the mean period from Step 1 of this
procedure (i.e., λ=H ¼ V s T m =H), and compute the associated
normalized value of PE . Kinematic interaction is anticipated
The frequency domain solution has the following five steps: to be significant if the wall under consideration lies near
1. Compute the Fourier transform of the free-field ground the fundamental-mode peak response region (i.e., λ=H ≈
displacement record ûg0 ðωÞ using a fast Fourier-transform 1.5 − 4), and small in regions of lower frequency
algorithm. (e.g., λ=H > ∼10).
2. Compute frequency-dependent values of the stiffness param- 5. Estimate ug0 so that the dimensionless wall force from Step 4
eters kiy , kiz , K y , K xx , K y emb , and K xx emb using Eqs. (7), (8), of this procedure can be dimensionalized. Ground motion am-
(9a), (9b), (10a), and (10b). Use typical protocols (NIST 2012) plitude ug0 should not be perceived as the peak ground displa-
for selecting representative shear moduli for use in these cement, but rather as a displacement associated with the most
expressions, including averaging nonuniform shear-wave energetic portion of the record. Until more detailed validation
velocities over appropriate depth ranges and using applicable exercises can be performed, ug0 should be taken as PGV=ωm ,
levels of modulus reduction for nonlinear problems (described where PGV is the peak ground velocity in the free field, and
further in a subsequent paragraph). Alternative values for ωm is the angular mean frequency corresponding to the mean
embedded foundation stiffness to those given in Eqs. (10a) period from Eq. (1), i.e., ωm ¼ 2π=T m . This recommendation
and (10b), as derived from site-specific and structure-specific is motivated by observations that energetic portions of the
analysis or from alternate solutions in the literature, can be ground motion spectrum are correlated with PGV (e.g., Akkar
readily incorporated by entering the computed values for and Özen 2005; Bommer and Alarcón 2006).
K y emb and K xx emb . This could be particularly important for 6. The total demand on the retaining wall is computed from PE
foundation geometries that are not well-approximated as plane and the resultant of the initial earth pressure, as in the FD
strain for a particular direction of shaking [e.g., rectangular procedure.
foundations, for which impedance solutions are available in Some important issues arise when selecting a representative
Gazetas (1983), Mylonakis et al. (2006), and NIST (2012)]. shear wave velocity using either the FD or SF solutions. First, shear
Material damping may also be incorporated through the use of wave velocity typically varies with depth due to pressure depend-
complex-valued shear moduli as noted previously. ence of soil shear modulus and age. For computing kiy and kiz , the
3. Compute the Fourier coefficients of the frequency-dependent time-averaged shear wave velocity (depth/travel time) for the depth
foundation input motions ûFIM ðωÞ and θ̂FIM ðωÞ using Eqs. (6a) interval from the ground surface to the bottom of the wall should be
and (6b). The term ûg0 ðωÞ is substituted for ug0 in Eqs. (6a) used. For computing base stiffness terms, the time-averaged shear
and (6b) for the frequency domain solution. wave velocity for the depth interval from z ¼ H to H þ B should
4. Compute the Fourier coefficients of the seismic earth pres- be used, until more detailed recommendations can be developed.
sure resultant, P̂E ðωÞ, using Eq. (5a). The terms ûg0 ðωÞ, Second, strong ground motion induces shear strains that are
ûFIM ðωÞ, and θ̂FIM ðωÞ are substituted for ug0, uFIM , and θFIM , large enough to reduce the secant shear modulus in accordance
respectively. with a modulus reduction curve. Failing to account for modulus

© ASCE 04015031-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


reduction may result in a significant overprediction of earth Sitar (2009, 2010) performed centrifuge modeling of embedded
pressure since the reduction in secant shear modulus reduces kiy , U-shaped walls, and concluded that M–O theory significantly over-
kiz , K y , and K xx . A site-specific ground response analysis is recom- predicts measured earth pressures. On the basis of their test results,
mended to obtain values of strain-compatible shear modulus they reported that dynamic earth pressures driving flexural de-
(and associated equivalent-linear V s ). An alternative crude ap- mands on the walls are negligible for peak horizontal surface
proach is to approximate the peak shear strain as PGV=V s . [Assum- accelerations less than 0.4g.
ing the standing wave field in Fig. 2 varies in time according
to ug ðz; tÞ ¼ ug0 · cosðkzÞ · eiωt , the ground surface velocity
is dug ð0; tÞ=dt ¼ u̇g0 ¼ i · ω · ug0 · eiωt and the shear strain is Ostadan (2005) Numerical Solution
dug =dz ¼ −k · ug0 · sinðkzÞ · eiωt . Therefore the strain field is Ostadan (2005) input six broadband earthquake motions, scaled to
dug =dz ¼ ðu̇g0 =V s Þ · i · sinðkzÞ, the amplitude of which is simply a common peak horizontal acceleration of 0.3g, to the base of an
PGV=V s .] The imaginary number indicates that shear strain is elastic soil layer with V s ¼ 305 m=s, H ¼ 9.14 m, mass density
90° out of phase with surface velocity. Furthermore, the maximum ρ ¼ 2.06 Mg=m3 , ν ¼ 1=3, and ξ ¼ 5%. The elastic layer rests
values of shear strain occur at the so-called nodes of the standing atop a rigid base. This elastic layer is the backfill behind a rigid
wave (i.e., at kz ¼ π=2 þ nπ, where n is an integer). For more wall also supported on the rigid base. The ground motions gener-
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

complicated conditions including soil layering and propagation ated substantial site response due both to the infinite impedance
of surface waves, shear strain has been found to range from 0.2 contrast (from the rigid base) and significant energy in the input
times to 1.7 times PGV=V s , with 1.0 being a commonly used value motions at the fundamental frequency of the backfill (where
for horizontal-component ground motions (Trifunac et al. 1996; λ=H ¼ 4).
Brandenberg et al. 2009), which provides an estimate of peak shear Five of the free-field surface motions were obtained from
strain consistent with the assumed shape of the soil displacement Ostadan (personal communication, 2013) and used to compute
profile. ug0 ðtÞ by double-integrating the surface accelerations in time.
Peak shear strain computed using these procedures can then be Those free-field motions were then applied using the proposed
converted to a representative uniform strain by multiplying the peak FD and SF solutions. Since the base of the wall was rigidly con-
shear strain by ðM − 1Þ=10, where M is moment magnitude (Idriss nected to the ground, only the stiffness term kiy is needed in the
and Sun 1991). The equivalent uniform shear strain would then be solution, and the frequency-dependent value was computed using
used to compute a value of G=Gmax from a selected modulus re- Eq. (7) with χy ¼ 1. Fig. 6(a) compares maximum earth pressures
duction curve, from which reduced values of G and V s can be ob-
over the wall height from the FD solution relative to those obtained
tained for use in the analysis. This equivalent-linear procedure
by Ostadan (2005) for two of the ground motions [three are omitted
neglects local strains imposed by the wall, and is reasonable for
for clarity in Fig. 6(a)]. Table 1 presents the resultants of these
cases involving free-field ground strains smaller than about 1%.
distributions. The resultant forces are in good agreement, with
However, the procedure may become erroneous at larger strains
errors ranging from −10 to þ12%.
corresponding to ground failure. Free-standing retaining walls that
In the SF solution, the surface displacement is computed as
rotate or translate significantly may mobilize such large shear
ug0 ¼ PGV · T=2π, where PGV is taken from ground-surface mo-
strains, but this will rarely be the case for stiff building basement
tions, and period T is taken as 4H=V s due to the strong impedance
walls.
contrast at the base of the soil layer. The agreement with the
The solution presented in this paper assumes perfect contact be-
Ostadan (2005) solution is reasonable, but not as good as the FD
tween the soil and the vertical walls. In reality, a gap might form in
solution, with errors ranging from −12 to þ57%. The Mononobe–
cohesive soils at this interface if PE is negative (i.e., the wall is
Okabe earth pressure resultant presented by Ostadan (2005),
moving away from the soil) and its absolute value is larger than
the initial earth pressure on the wall. Gapping may theoretically 160 kN=m for all of ground motions, underpredicts the earth pres-
cause pounding and additional stresses on the wall beyond those sures in every case.
considered in this paper. However, it is likely that peak earth pres- The conditions considered by Ostadan (2005) are nearly optimal
sures will occur when PE is positive (i.e., when the free-field soil for generating large kinematic wall pressures (i.e., λ=H ¼ 4, asso-
moves towards the wall), which is considered in the present analy- ciated with first mode response of the backfill, lies near the peaks
sis. The efficacy of the proposed procedure is demonstrated in the of the curves in Fig. 4). Such conditions cause the mobilized earth
subsequent section and will be tested further over time as additional pressures to exceed those from the M–O theory. The Ostadan
experimental data become available. (2005) results are broadly consistent with previous findings by
Arias et al. (1981) and Veletsos and Younan (1994) obtained by
analytical closed-form solutions for similar configurations.
Comparison to Experimental-Based and
Simulation-Based Results in Literature Al Atik and Sitar (2009, 2010) Experimental Results
In this section, two prior studies that reached strongly divergent Al Atik and Sitar (2009, 2010) performed centrifuge experiments
conclusions about the levels of seismic earth pressures acting on on relatively rigid and flexible U-shaped walls with prototype di-
retaining walls are interpreted using the proposed methodology. mensions of H ¼ 6.5 m and B ¼ 5.3 m embedded in a profile of
In the first study, Ostadan (2005) performed elastic wave propaga- medium dense sand with thickness D ¼ 19 m, and γ ¼ 17 kN=m3 .
tion analysis using a numerical finite-element code (SASSI; Lysmer The average small-strain shear wave velocities given by Al Atik
et al. 1999) to investigate the kinematic interaction between free- and Sitar (2009, 2010) were V s ¼ 170 m=s behind the walls
field site response and a massless embedded structure connected to and V s ¼ 260 m=s for the depth interval from the base of the wall
a rigid base and fixed against rotation. Ostadan (2005) concluded to the essentially rigid base of the container. The FD and SF so-
that M–O earth pressure theory significantly underpredicts the mo- lutions are compared with results of experiments performed
bilized earth pressures by factors ranging from 2 to 4 depending on using motions denoted Loma Prieta SC1, Loma Prieta SC2, and
ground motion characteristics. In the second study, Al Atik and Kobe PI2.

© ASCE 04015031-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Maximum seismic earth pressure increments computed by the proposed kinematic methodology compared with (a) numerical solutions by
Ostadan (2005); (b) experimental results by Al Atik and Sitar (2009, 2010)

For these experiments, ug0 was obtained by digitizing and the predicted distributions being approximately zero at the base
double-integrating in time the plots of free-field surface accelera- of the wall and having their maximum at the ground surface. This
tion presented by Al Atik and Sitar (2009). These motions induced mismatch may result in part from the assumption of depth-invariant
nonlinear response in the sand, and measured shear strains and kiy , whereas the shear modulus of sand in the centrifuge models
the interpreted modulus reduction (G=Gmax ) curve by Al Atik and increases with depth. A more robust solution would utilize kiy val-
Sitar (2009, 2010) were used to estimate representative values of ues that increase with depth in accordance with the variation in soil
G=Gmax ¼ 0.28, 0.25, and 0.10 for the SC1, SC2, and PI2 ground shear modulus, combined with a site response study that captures
motions, respectively. Comparisons between computed (FD solu- the influence of these variations on the free-field displacement
tion) and measured maximum earth pressures for the three digitized profile. The authors lacked the required data to perform such an
ground motions are shown in Fig. 6 for SC2 and PI2 (SC1 omitted analysis. The modulus reduction was an important part of this
for clarity). Resultant forces for all three motions are shown in analysis; if taken as unity (linear soil) earth pressures are signifi-
Table 2. Resultant force errors range from −7 to þ23% for the cantly overpredicted.
FD solution and from þ6 to þ23% for the SF solution. Although Mononobe–Okabe earth pressures presented by Al Atik and
the earth pressure resultants are predicted quite well, the shape of Sitar (2009, 2010) were computed using the ground surface PGA
the pressure distributions differ significantly, with the reported dis- and 0.65PGA. For consistency with the Ostadan (2005) compari-
tributions from measurements increasing linearly with depth and sons, the PGA-based M–O estimates are presented here. As shown

Table 2. Resultant of Seismic Earth Pressure Increments from Al Atik and


Table 1. Resultants of Seismic Earth Pressure Increments from Ostadan Sitar (2009), the Mononobe–Okabe Solution, and the Proposed Kinematic
(2005), the Mononobe–Okabe Solution, and the Proposed Kinematic Methodology
Methodology
Earth pressure resultant, PE (kN=m)
Earth pressure resultant, PE (kN=m)
Al Atik Frequency- Single- Mononobe–
Frequency- Single- Mononobe– and Sitar domain frequency Okabe
Ostadan domain frequency Okabe Ground motion (2009) solutiona solutiona solutiona
Ground motion (2005) solutiona solutiona solutiona
Loma Prieta SC1 90 110 95 180
Loma Prieta 414 415 487 160 þ23 þ6 þ100
(þ0%) (þ18%) (−61%) Kobe PI2 146 164 180 ∞
ATC 368 341 461 160 þ13 þ23 þ∞b
(−7%) (þ25%) (−57%) Loma Prieta SC2 101 94 121 235
RG1.60 478 451 588 160 −7 þ20 þ132
(−6%) (þ23%) (−67%) a
The second value in each cell, i.e., after the line break, is the percent error.
EUS distant 405 362 637 160 b
The M–O prediction of infinite earth pressure is caused by the inertial
(−11%) (þ57%) (−60%)
force exceeding the shear strength of the sand at the base of the wall,
EUS local 179 201 158 160
and is a well-recognized unrealistic artifact that makes the M–O theory
(þ12%) (−12%) (−11%)
difficult to apply in practice for sites with very strong design ground
a
The second value in each cell, i.e., after the line break, is the percent error. motions.

© ASCE 04015031-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


in Table 2, the M–O pressure resultants significantly exceed the for design application. Nevertheless, the proposed approach produ-
measurements. It is helpful to visualize these results relative to ces estimates of seismic earth pressures that are significantly more
the diagrams in Fig. 4. If the frequency content of the motions accurate than M–O theory.
in the centrifuge model are assumed to be dominated by site re- Numerical simulations are warranted for cases where the as-
sponse above the essentially rigid base of the container, then sumptions associated with the proposed method are expected to
λ ¼ 4D, which produces λ=H ¼ 12. This is well to the right of the produce unacceptably large errors. Seismic earth pressures from
peak, and therefore anticipated soil pressures from kinematic inter- inertial interaction should also be considered in general application,
action are quite small. Not surprisingly, those pressures fall below and may be the principal source of seismic earth pressures when
the range of M–O predictions. kinematic interaction is insignificant. Inertial demands have dif-
Fig. 6 and Table 2 compare results from the proposed analysis ferent origins, and as such, may be out of phase with kinematic
with maximum kinematic earth pressure increments presented by demands. Inertia demands should be evaluated separately using
Al Atik and Sitar (2009, 2010), i.e., total earth pressure minus ini- a procedure like that shown in Fig. 1(b) and described in detail
tial static earth pressure minus the component from inertia of the elsewhere (e.g., NIST 2012).
wall mass. However, Al Atik and Sitar (2009, 2010) indicate that
the peak bending moments in the walls arose from a combination
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

of kinematic and inertia loading, and peak moments were out-of- Acknowledgments
phase with peak kinematic earth pressures. The evaluation of these
inertial effects is a straightforward extension of the proposed meth- The authors thank Farhang Ostadan for sharing the ground motion
odology, but is not considered in this paper for brevity and because data utilized in a 2005 paper, and the two anonymous reviewers
the authors lacked the required data. whose comments helped improve the paper.

Effect of Dynamic Modifier on Lateral Wall–Soil


Stiffness Terms Supplemental Data
Calculations of PE presented previously utilized frequency- Fig. S1 and Eqs. S1–S17 are available online in the ASCE Library
dependent stiffness terms Eqs. (7) and (8) for both the FD and (www.ascelibrary.org).
SF solutions. The calculations were repeated omitting the dynamic
component (i.e., setting ω ¼ 0). Setting the frequency modifiers to
unity increased the computed earth pressures by about 15–20% for References
the FD solution for both the Ostadan (2005) and Al Atik and Sitar
(2009, 2010) cases. This generally increases model misfit to the Akkar, S., and Özen, Ö. (2005). “Effect of peak ground velocity on defor-
data from the literature. Using the SF solution, comparable pressure mation demands for SDOF systems.” Earthquake Eng. Struct. Dyn.,
increases for the Al Atik and Sitar (2009, 2010) case are observed, 34(13), 1551–1571.
but >200% increases are observed for the Ostadan (2005) case. Al Atik, L., and Sitar, N. (2009). “Experimental and analytical study of
On the basis of these comparisons, application of the frequency- the seismic performance of retaining structures.” Rep. No. Pacific
dependent terms in Eqs. (7) and (8) is recommended when the Engineering Earthquake Engineering Research (PEER)-2008/104,
interaction effects are strong (i.e., near the peak of the transfer PEER Center, Univ. of California, Berkeley, CA.
functions in Fig. 4, or λ=H ≈ 1.5 − 5.0. Otherwise, for the Al Atik, L., and Sitar, N. (2010). “Seismic earth pressures on cantilever
retaining structures.” J. Geotech. Geoenviron. Eng., 10.1061/(ASCE)
common case of λ=H > 5, implementation of the dynamic modi-
GT.1943-5606.0000351, 1324–1333.
fier appears to be helpful but not essential.
Arias, A., Sanchez-Sesma, F. J., and Ovando-Shelley, E. (1981). “A sim-
plified elastic model for seismic analysis of earth-retaining structures
with limited displacements.” Proc., Int. Conf. on Recent Advances in
Recommendations and Conclusions Geotechnical Earthquake Engineering and Dynamics, S. Prakash, ed.,
Vol. 1, Univ. of Missouri, Rolla, MO, 235–240.
We present a kinematic soil-structure interaction approach that pro-
Bommer, J. J., and Alarcón, J. E. (2006). “The prediction and use of peak
vides a unifying framework to explain the lower-than-M–O seismic ground velocity.” J. Earthquake Eng., 10(1), 1–17.
earth pressure increments observed by Al-Atik and Sitar (2009, Brandenberg, S. J., Coe, J., Nigbor, R. L., and Tanksley, K. (2009). “Differ-
2010) and the higher-than-M–O pressure increments computed by ent approaches for estimating ground strains from pile driving vibra-
Ostadan (2005), Veletsos and Younan (1994), and others. The ap- tions at a buried archeological site.” J. Geotech. Geoenviron. Eng.,
proach is admittedly simplified in several respects; in particular, 10.1061/(ASCE)GT.1943-5606.0000031, 1101–1112.
(1) the effects of wall and foundation inertia are not considered Chen, W. F. (1975). Limit analysis and soil plasticity. Developments in
(consistent with a kinematic assumption), (2) the Winkler as- geotechnical engineering, Elsevier, Amsterdam, Netherlands.
sumption is utilized, (3) the single-frequency solution significantly Chen, W. F., and Liu, X. L. (1990). Limit analysis in soil mechanics,
simplifies the broadband ground motion driving the kinematic de- Elsevier, Amsterdam, Netherlands.
mands, (4) soil nonlinearity can only be indirectly included using Elsabee, F., and Morray, J. P. (1977). Dynamic behavior of embedded
an equivalent-linear approximation, and (5) potential impacts of foundations, Massachusetts Institute of Technology, Cambridge, MA.
alternate initial gravity-induced stress conditions (e.g., active and Gazetas, G. (1983). “Analysis of machine vibrations: State of the art.” Soil
Dyn. Earthquake Eng., 2(1), 2–42.
at-rest) on the seismic earth pressure increment are not considered.
Gazetas, G. (1991). “Foundation vibrations.” Chapter 15, Foundation
Despite those caveats, the approach is physically sound and pro-
engineering handbook, 2nd Ed., H.-Y. Fang, ed., Chapman and Hall,
vides a clear basis for understanding the factors driving seismic New York.
earth pressures for many practical retaining wall configurations. Gazetas, G., and Roesset, J. M. (1976). “Forced vibrations of strip footings
Additional experimental observations and numerical simulations on layered soils.” Proc., National Structural Engineering Conf., Vol. 1,
are needed to validate the procedure for ranges of ground motion W. E. Saul and A. H. Peyrol, eds., ASCE, Reston, VA, 115–131.
frequencies and wall configurations, evaluate the relative contribu- Idriss, I. M., and Sun, J. I. (1991). User’s manual for SHAKE91, Center for
tions of inertial effects, and to formulate detailed recommendations Geotechnical Modeling, Univ. of California, Davis, CA.

© ASCE 04015031-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.


Jakub, M., and Roesset, J. M. (1977). “Nonlinear stiffness of foundations.” NIST. (2012). “Soil-structure interaction for building structures.” Rep. NIST
Research Rep. R77-35, Massachusetts Institute of Technology, Grant/Contract Rep. (GCR) 12-917-21, J. P. Stewart, ed., Gaithersburg,
Cambridge, MA. MD.
Kausel, E., Whitman, A., Murray, J., and Elsabee, F. (1978). “The spring Okabe, S. (1924). “General theory of earth pressure and seismic stability of
method for embedded foundations.” Nucl. Eng. Des., 48(2–3), retaining wall and dam.” J. Jpn. Soc. Civ. Eng., 12(4), 34–41.
377–392. Ostadan, F. (2005). “Seismic soil pressure for building walls—An updated
Kloukinas, P., Langoussis, M., and Mylonakis, G. (2012). “Simple approach.” Soil Dyn. Earthquake Eng., 25(7–10), 785–793.
wave solution for seismic earth pressures on non-yielding walls.” J. Pais, A., and Kausel, E. (1988). “Approximate formulas for dynamic stiff-
Geotech. Geoenviron. Eng., 10.1061/(ASCE)GT.1943-5606.0000721, nesses of rigid foundations.” Soil Dyn. Earthquake Eng., 7(4), 213–227.
1514–1519. Rathje, E. M., Faraj, F., Russell, S., and Bray, J. D. (2004). “Empirical
Lew, M., Sitar, N., and Al-Atik, L. (2010). “Seismic earth pressures: Fact or relationships for frequency content parameters of earthquake ground
fiction.” Proc., Earth Retention Conf., R. Finno, Y. M. A. Hashash and motions.” Earthquake Spec., 20(1), 119–144.
P. Arduino, eds., ASCE, Reston, VA. Seed, H. B., and Whitman, R. V. (1970). “Design of earth retaining struc-
tures for dynamic loads.” Proc., ASCE Specialty Conf. on Lateral
Li, X. (1999). “Dynamic analysis of rigid walls considering flexible foun-
Stresses in the Ground and Design of Earth Retaining Structures, Vol. 1,
dation.” J. Geotech. Geoenviron. Eng., 10.1061/(ASCE)1090-0241
Reston, VA, 103–147.
(1999)125:9(803), 803–806.
Steedman, R. S., and Zeng, X. (1990). “The influence of phase on
Downloaded from ascelibrary.org by Xavier Vera on 07/02/15. Copyright ASCE. For personal use only; all rights reserved.

Lysmer, J., Ostadan, F., and Chin, C. (1999). “SASSI2000 theoretical


the calculation of pseudo-static earth pressure on a retaining wall.”
manual.” Rep. Prepared for the Geotechnical Engineering Division,
Geotechnique, 40(1), 103–112.
Civil Engineering Dept., Univ. of California, Berkeley, CA.
Trifunac, M. D., Todorovska, M. I., and Ivanović, S. S. (1996). “Peak
Mononobe, N., and Matsuo, M. (1929). “On the determination of earth velocities and peak surface strains during Northridge, California, earth-
pressures during earthquakes.” Proc., World Engineering Congress, quake of 17 January 1994.” Soil Dyn. Earthquake Eng., 15(5), 301–310.
Engineering Society of Japan, Tokyo, 179–187. Varun, Assimaki, D., and Gazetas, G. (2009). “A simplified model for lat-
Mushkelishvili, N. I. (1963). Some basic problems of the mathematical eral response of large diameter caisson foundations—Linear elastic for-
theory of elasticity, P. Noordhoff, Groningen, Netherlands. mulation.” Soil Dyn. Earthquake Eng., 29, 268–291.
Mylonakis, G., Nikolaou, S., and Gazetas, G. (2006). “Footings under Veletsos, A. S., and Prasad, A. M. (1989). “Seismic interaction of structures
seismic loading: Analysis and design issues with emphasis on bridge and soils: Stochastic approach.” J. Struct. Eng., 10.1061/(ASCE)0733
foundations.” Soil Dyn. Earthquake Eng., 26(9), 824–853. -9445(1989)115:4(935), 935–956.
NCHRP (National Cooperative Highway Research Program). (2008). Veletsos, A. S., and Younan, A. H. (1994). “Dynamic soil pressures on rigid
“Seismic analysis and design of retaining walls, buried structures, retaining walls.” Earthquake Eng. Struct. Dyn., 23(3), 275–301.
slopes, and embankments.” Rep. 611, D. G. Anderson, G. R. Wood, J. H. (1973). “Earthquake induced soil pressures on structures.”
Martin, I. P. Lam, and J. N. Wang, eds., National Academies, Rep. No. Earthquake Engineering Research Laboratory (EERL) 73-05,
Washington, DC. California Institute of Technology, Pasadena, CA.

© ASCE 04015031-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2015.141.

You might also like