You are on page 1of 12

Coupled Seepage-Deformation Model for Predicting

Pore-Water Pressure Response during Tsunami Loading


Abbas Abdollahi 1 and H. Benjamin Mason 2
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Tsunamis induce pore-water pressures in coastal soil beds, and understanding how tsunami-induced excess pore-water pressure
gradients form is not straightforward given the loading timescale and seepage effects. This study develops a coupled seepage-deformation
model to describe tsunami-induced pore-water pressure response in soil beds. The coupled seepage-deformation model is implemented using
the finite-difference method, and numerical experimentation is performed on saturated soil beds using hypothetical tsunamis. The results
imply the dependence of the tsunami-induced excess pore-water pressure head gradient on the tsunami water’s weight and compressibility of
the pore water. The results imply the importance of modeling tsunami-induced deformation, the partial drainage condition, and pore-fluid
condition when considering tsunami-induced pore-water pressures in soil beds. DOI: 10.1061/(ASCE)GT.1943-5606.0002012. © 2019
American Society of Civil Engineers.
Author keywords: Fluidization; Finite-difference method; Liquefaction; Partial drainage; Pore-water pressure; Seepage-deformation
model; Tsunami; Wave loading; Richard’s equation.

Introduction Researchers also investigated pore-water pressure response


caused by water waves (e.g., Putnam 1949; Mei and Foda 1981;
During tsunami run-up, pore-water pressures in soil beds increase Zen and Yamazaki 1990a, b; Sakai et al. 1991; Gratiot and
as the soil bed pressurizes. Then, during tsunami drawdown, the Mory 2000; Chowdhury et al. 2006; Liu et al. 2007; Young
soil bed depressurizes. The depressurization causes a gradient of et al. 2009; Sumer et al. 2011; Merxhani and Liang 2012; Kirca
excess pore-water pressure head, ∂h=∂z, which results in upward et al. 2013; Sumer 2014). Yet, the types of water waves that the
seepage of pore water. If the upward seepage becomes fast enough, preceding researchers investigated have a smaller period than
then fluidization of the soil bed occurs (Lowe 1975). In particular, tsunamis, so the models are not appropriate. In addition, the afore-
fluidization, or tsunami-induced liquefaction, occurs when ∂h=∂z mentioned researchers neglected the compressibility of the solid
becomes greater than or equal to ðρsat =ρÞ − 1; where ρsat is the bulk constituent relative to the fluid bulk modulus. Herein, the focus
mass density of a saturated soil bed and ρ is the mass density of is on tsunami waves as well as keeping the model general (e.g., sen-
water. Tsunami-induced liquefaction is different from earthquake-
sitivity of the results to tsunami characteristics and compressibility
induced liquefaction, which is caused by the residual build-up of
of pore fluid). Tsunami drawdown on a soil bed is also like rapid
excess pore-water pressure during cyclic shaking and initiates at
drawdown of water in embankment dams (e.g., Bishop 1954; Lowe
some depth below the ground surface in relatively loose, coarse-
and Karafiath 1960; Duncan et al. 1990; Fredlund et al. 2011), but
grained soils (Seed 1979; Mason and Yeh 2016).
the timescales are different. The rapid drawdown of water in em-
Yeh and Mason (2014) investigated tsunami-induced pore-water
pressure in elastic, semi-infinite sand beds using an uncoupled bankment dams occurs on the order of days; in contrast, tsunami
assumption. That is, the pore-water pressure in the soil bed only run-up and drawdown occurs on the order of minutes.
changes when the overlying tsunami water seeps into the soil bed. The preceding discussion shows that models developed to esti-
The uncoupled assumption is a limitation, and coupled seepage- mate tsunami-induced pore-water pressures are rare. One reason for
deformation models relax the limitations (Biot 1941, 1956). the paucity of appropriate models is the uniqueness of tsunami
Loading-specific seepage-deformation models couple pore-water loading (Bernard and Robinson 2009). In particular, tsunamis have
pressure and deformation responses in soil beds to different loading large amplitudes and long periods. A typical tsunami height is on
types (e.g., Ochoa-Tapia and Whitaker 1995; Zienkiewicz et al. the order of one to tens of meters and inundates several hundred
1999; Discacciati et al. 2002; Goyeau et al. 2003; Balhoff et al. meters or more, and tsunamis have depth-averaged flow velocities
2008; Mosthaf et al. 2011; Chen and Wang 2014; Bukač et al. 2015; of 10 m=s or higher. As a result, tsunami loading is rapid and tran-
Mayr et al. 2015). These aforementioned studies are not applicable sient. Moreover, tsunami loading creates partial drainage, imparts
to tsunami loading without significant changes. large total stresses, and causes seepage.
The authors present a coupled model to predict tsunami-induced
1
Geotechnical Engineer, AECOM Technical Services, Inc., 2020 L St., pore-water pressures in soil beds. Partial drainage is incorporated
Sacramento, CA 95811 (corresponding author). Email: abbas.abdollahie@ by classifying the excess pore-water pressure in two parts: (1) the
gmail.com; abbas.abdollahi@aecom.com undrained part, which corresponds to the pore-water pressure in-
2
Associate Professor, School of Civil and Construction Engineering, duced by the weight of the overlying tsunami water; and (2) the
Oregon State Univ., 450 Learning Innovation Center, Corvallis, OR 97331.
drained part, which corresponds to the pore-water pressure induced
Note. This manuscript was submitted on November 2, 2017; approved
on August 20, 2018; published online on January 7, 2019. Discussion per- by seepage. The results of numerical experiments are presented to
iod open until June 7, 2019; separate discussions must be submitted for confirm the proposed model. The implicit goal is to keep the
individual papers. This paper is part of the Journal of Geotechnical model simple to elicit understanding of underlying physical mech-
and Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. anisms and encourage its use.

© ASCE 04019002-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


Table 1. Fundamental assumptions required for the proposed seepage-deformation model
Number Assumption Comments
1 Cyclic deformation within the soil bed The assumption is acceptable because tsunami wave periods are large, and consequently, the pore-water
is neglected (i.e., residual liquefaction pressure induced by cyclic deformation within the soil bed is small. Neglecting cyclic deformation
is neglected). means also neglecting the reorganization of soil particles at the soil bed’s surface, which can spatially
and temporally change the near-surface porosity of the soil bed and affect the tsunami-induced
pore-water pressure response.
2 The soil bed is fully saturated. An appropriate assumption when investigating nearshore, gently sloping soil beds. The assumption
implies that the degree of saturation, S, in Eq. (3) = 1.
3 The strain condition is uniaxial, where An appropriate simplification because the tsunami-induced deformation in the x- and y-directions is
the axial direction is the z-direction. negligible compared with the corresponding deformation in the z-direction. The assumption implies that
one neglects lateral stresses imposed by the tsunami flow as well as lateral pore-water pressure gradients.
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

4 At each time step, the soil bed is in The pore-water pressure induced by the weight of the overlying tsunami water is applied to the entire soil
equilibrium and the loading condition column instantaneously, and the tsunami loading occurs relatively rapidly.
is undrained.
5 Absolute permeability, kab, is constant. Absolute permeability changes as a function of pore size, porosity, and the tortuosity of the porous
medium (Bear 1988); however, porosity changes within a soil bed are typically small for tsunami loading
(porosity is expected to have a vertical variation near the ground surface, but this study focuses on
average porosity changes within the entire soil bed).
6 Darcy’s law is valid. An appropriate assumption when the pore water is incompressible; however, the assumption becomes
less appropriate as the pore water becomes more compressible (Hubbert 1940). Appendix II provides a
further discussion.

Proposed Seepage-Deformation Model The proposed deformation model in Eq. (2) must be coupled
with a seepage model. The differential equation governing the
The proposed seepage-deformation model is based on a deforma- flow in soil beds (neglecting source or sink terms) can be written
tion model advanced by Fredlund and Morgenstern (1976) and (Richards 1931)
Fredlund et al. (2012), which is ultimately based on the deforma-  
tion model by Biot (1941) and is given as follows: ∂ ρkab kr  
ðρϕSÞ − ∇ · ∇pw − ρg∇D ¼ 0 ð3Þ
∂t μ
ð1 þ νÞ
δϵv ¼ mv ½δðσ − pa Þ þ ½δðpa − pw Þ ð1Þ As before, Appendix III contains modeling parameters, and
Hð1 − νÞ
Table 1 lists major assumptions required to use Eq. (3); in addition,
D = depth-of-interest within the porous medium (i.e., for vertical
The Fredlund et al. (2012) deformation model was selected as a
flow, D ¼ z); and ∇ = gradient operator.
framework for the proposed seepage-deformation model because it
The seepage model [Eq. (3)] and deformation model [Eq. (2)]
is a two-state variable deformation model. That is, it relates the in-
are coupled by the porosity, φ. In other words, Eq. (3) is a function
crement of volumetric strain, εv , to the increments of net stress,
of φ, which is calculated at each time step using Eq. (2). The
σ − pa , and matric suction, pa − pw , where σ is the total vertical
coupling requires a fundamental assumption, which is not listed
stress, pa is the pore-air pressure, and pw is the total pore-water
in Table 1: the increment of volumetric strain, δεv [calculated with
pressure. In addition, the Fredlund et al. (2012) deformation model
the deformation model Eq. (2)] is equal to the increment of poros-
is easy to implement and helps elicit understanding. Appendix III
ity, δφ [input into the seepage model Eq. (3)]. The difference be-
provides the definitions or values of additional modeling parame-
tween the assumption made here (i.e., δφ ¼ δεv ) and more realistic
ters necessary for understanding and using Eq. (1), and Table 1
condition is that this study assumes the variation of volume change
presents the primary assumptions used to define the proposed
is relative to the initial volume of the soil bed and not the instanta-
seepage-deformation model. The porosity of the soil bed, φ,
neous volume of the soil bed. The assumption that δφ ¼ δεv
changes as a function of tsunami-induced deformation, and an
implies that δεv ¼ δe=ð1 þ eÞ, where δe is the increment of void
initial soil bed porosity, φ0 , of 0.4 for beach sand was assumed
ratio. Accordingly, the δφ ¼ δεv assumption is consistent with
(Santamarina et al. 2001).
the typical assumption used in soil mechanics when examining
Eq. (1) was modified using Appendix III and Table 1, which
the volume change behavior of soils (Lambe and Whitman 1969).
leads to
 
Bð1 þ ν u Þ
δϵv ¼ mv δσ − αδpw ð2Þ Numerical Experimentation
3ð1 − ν u Þ

Numerical Formulation
Eq. (2) is the final functional form of the proposed deformation
model for a fully saturated soil bed with all the simplifying assump- Numerical solutions for Eq. (3) are pervasive (e.g., Zarba et al.
tions and appropriate modifications for loading of a porous medium 1990; Woodward and Dawson 2000), and the fully implicit cell-
by an overlying fluid. The process of reducing Eqs. (1) and (2) centered finite difference approach was used, as implemented in
is explained in Appendix I (Abdollahi 2017). As a brief conceptual MATLAB R2016b (version 9.1) (Peaceman 1977; Helmig 1997;
aside, if one assumes a typical consolidation test on a soil Peszyńska et al. 2002; Peszyńska and Yi 2008). When solving
sample, then B ¼ 1 and υu ¼ 0.5, which leads to mv ¼ δεv =δσ 0 , Eq. (3), it was assumed that the pore-water pressure is the primary
i.e., the definition of mv for consolidation testing (Lambe and unknown. During the first step, the number of grids, time step, time
Whitman 1969). range, method of averaging, type of boundary conditions, and

© ASCE 04019002-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


Q1 Q2 Q3 Qj Q j+1 Q m Q m+1 Table 2. Numerical experimentation cases with defined nomenclature
Case Model B α νu Type
0 p pw2 pwj L S-LE-B(0) LE 0 1 0.5 —
w1 Pwm+1
S-LE-B(0.5) LE 0.5 1 0.5 —
Fig. 1. Graphical depiction of the finite-difference cell-centered spatial S-LE-B(0.6) LE 0.6 1 0.5 —
domain. S-LE-B(0.7) LE 0.7 1 0.5 —
S-LE-B(0.8) LE 0.8 1 0.5 —
S-LE-B(0.9) LE 0.9 1 0.5 —
S-LE-B(0.95) LE 0.95 1 0.5 —
S-LE-PW LE — — — Deaired
values of the boundary conditions are selected, and then the initial
S-LE-UW LE — — — Aired
conditions are defined. At the initial time, the soil properties are S-NI-B(0) NI 0 1 0.5 —
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

defined and computed. To perform the computations, the spatial S-NI-B(0.5) NI 0.5 1 0.5 —
domain must be defined, and it is discretized by a uniform grid S-NI-B(0.8) NI 0.8 1 0.5 —
of size Δz, with cell centers denoted by zj (Fig. 1). The subscript S-NI-B(0.95) NI 0.95 1 0.5 —
j defines the assigned cell number and is j ¼ 1; : : : ; mz. The time S-NI-PW NI — — — Deaired
domain is also discretized into subintervals of variable size ΔT n , S-NI-UW NI — — — Aired
where n ¼ 1; : : : ; N, where N is the total number of time steps. Note: In cases where the B value was not selected manually, B, α, and ν u
After the input data are defined, the computations are performed are estimated from poroelasticity.
for the defined time range. At the initial time step, the guess un-
knowns are equal to the values from the initial conditions. For sub-
sequent time steps, the guess unknowns are equal to the guess
2 2
unknowns from the previous time step. Based on the updated guess Tð1Þ ¼ Δz Tðmz þ 1Þ ¼ Δz ð8Þ
1 1 mz
unknowns from the previous time step or the previous iteration, the k1 kmz
hydraulic conductivity, pore-water pressure, and porosity—as well
as all the respective Jacobians—are computed. Then, the entire Based on the preceding discussion, at each iteration, i, the
matrix of Jacobians and matrix of residuals are calculated. The fundamental computation is
Jacobians, J, are computed analytically as follows:
Ji δi ¼ −Ri ð9Þ
 
∂Rj ∂Rj ∂Rj
J¼ ; ; ð4Þ which is solved for δ, and
∂pw;j ∂pw;jþ1 ∂pw;j−1
piw ¼ δ i þ pi−1
w ð10Þ
where Rj = residual value, which is calculated
When performing the preceding computations, the harmonic
1 h n i average was used to define the values of absolute permeability
ðρj ϕj Sj Þ − ðρj ϕj Sj Þ
n n−1
Rj ¼ − Qjþ1 − Qnj−1 ¼ 0 ð5Þ and dynamic viscosity for each cell at each iteration during each
Δt Δzj 2 2
time step. Similarly, arithmetic averaging was used to estimate
the average density and permeability for each cell at each iteration
where Q is calculated during each time step.

Qnjþ1 ¼ ρjþ12 krjþ12 T jþ12 ½pnwjþ1 − pnwj − pwjþ12 gðDjþ1 − Dj Þ ð6aÞ


2 Experimental Plan
This study focuses on very fine, poorly graded beach sands with a
Qnj−1 ¼ ρj−12 krj−12 T j−12 ½pnwj − pnwj−1 − pwj−12 gðDj − Dj−1 Þ ð6bÞ specific gravity of the soil grains, Gs , of 2.65. Appendix III con-
2
tains the important modeling parameters, and a constant depth to
and the transmissibility functions, T, are computed bedrock of 10 m was assumed. For the given soil bed, tsunami-
induced liquefaction occurs, roughly speaking, when ∂h=∂z ≥ 1,
2 and enhanced scour occurs when ∂h=∂z ≥ 0.5.
T j12 ¼ Δz Δzj1
ð7Þ Table 2 presents the performed numerical experiments and
kj
j
þ kj1 the nomenclature. The first letter, S, indicates the overarching
assumption that the soil bed is fully saturated. The second two let-
where k ¼ kab =μ. ters correspond to the soil stiffness model: linear (LE) versus non-
Two types of boundary conditions are considered: Neumann and linear (NI). The pore-water condition is given as deaired (PW) or
Dirichlet. For the Neumann boundary condition, the pore-water aired (UW), or for the cases where the B value is systematically
flow at the boundary is specified. Herein, no flow is allowed at changed, B is used followed by the specific B value in parentheses.
the impermeable boundary, so ∂pw =∂z at the impermeable layer The hypothetical tsunami is generated using the solution to the
is zero. For the Dirichlet boundary condition, the value of the nonlinear shallow-water wave equations given by Carrier et al.
pore-water pressure is defined at the boundary. Herein, the pore- (2003), which were implemented in MATLAB. Fig. 2 shows the
water pressure (in terms of head) at the ground surface is equal tsunami flow height time series at the shoreline for the hypothetical
to the height of the tsunami. Notably, the aforementioned Jacobian tsunami with a length scale of 40 km and beach slope of 1/100 (the
and residual matrices do not include the effects of the boundary initial wave shape was assumed to be Gaussian with an initial
conditions. Therefore, the Jacobians and residual values at the velocity of zero). The tsunami run-up occurs in about 25 min
boundaries are calculated and added to the matrices at the appro- and the tsunami drawdown occurs in about 7 min. The tsunami flow
priate indices, more specifically, at the boundaries of (1) and height time series was smoothed with a median filter, but it still has
(mz þ 1) some minor irregularities, which may result in minor irregularities

© ASCE 04019002-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Variation of a typical tsunami height with time in the shoreline


for (a) dimensionless wave height and time; and (b) dimensioned wave
height and time. The tsunami wave data are from the Carrier et al.
(2003) model. The slope of the beach is 1/100 and the length
scale, L, is 40 km. The dimensioned time series was developed from
the dimensionless time series using scaling factors given by Carrier
et al. (2003).

of the results of the numerical experimentation when the pore-water


pressure head gradient is investigated. These minor irregularities do Fig. 3. Convergence in terms of total pore-water pressure profiles for
not affect the overall trends. The hypothetical tsunami is similar to depths between 0 and 3 m for times of (a) 10 min; (b) 20 min;
realistic tsunamis measured in the field (e.g., Fritz et al. 2012; (c) 28 min; and (d) 29 min.
Koshimura and Hayashi 2012; Yeh and Sato 2016).

Convergence Analysis Fig. 5(b), which shows the excess pore-water pressure head gra-
A convergence analysis was performed as shown in Figs. 3 and 4. dient, ∂h=∂z, versus depth for different times. Recall that positive
Four convergence analyses were performed in total, and the values of ∂h=∂z indicates upward flow; accordingly, downward
relevant spatial and temporal parameters are given in Table 3. Con- flow occurs during tsunami run-up, and upward flow occurs during
vergence was achieved in terms of total pore-water pressure, pw , tsunami drawdown, which is observed in Fig. 5(b). The depth pro-
and excess pore-water pressure head gradient, ∂h=∂z, versus depth. files of ∂h=∂z are nearly linear, with the value being zero at the
Based on the convergence analysis, cell dimensions and time steps bedrock (i.e., Neumann boundary condition) and the largest value
of Δz ¼ 0.1 m and Δt ¼ 0.0639 s were chosen to perform the at the ground surface. The largest value of ∂h=∂z occurs at the end
numerical experiments. of drawdown, as expected, but it is not large enough to cause sig-
nificant sediment instability.
Fig. 6 shows ∂h=∂z versus time at the ground surface, 1 m be-
Results low the ground surface, and 5 m below the ground surface. The
Figs. 5 and 6 show the results of the hypothetical tsunami ∂h=∂z values calculated at the ground surface and at a depth of
[Fig. 2(b)] inundating the saturated soil model with a linear soil 1 m below the ground surface are nearly identical, which is indica-
constitutive model and aired water as the pore fluid (i.e., Case tive of the ∂h=∂z versus depth profiles being linear, as previously
S-LE-UW in Table 2). Fig. 5(a) shows the total pore-water pressure, discussed. In addition, Fig. 6 shows that the increase in ∂h=∂z is
pw , versus depth during tsunami run-up and drawdown. Recall more significant during the tsunami drawdown period than in the
from Fig. 2(b) that 20 min is during the middle of the tsunami tsunami run-up period, which indicates that the tsunami drawdown
run-up, 25 min is near the start of tsunami drawdown, and 30 min occurs relatively rapidly compared with the tsunami run-up, as
is near the end of tsunami drawdown. Fig. 5(a) shows that the shown in Fig. 2.
total pore-water pressure increases as the tsunami run-up is occur- A normalized bar chart was created to compare multiple cases.
ring and decreases relatively rapidly as the tsunami drawdown is The bar chart (Fig. 7) shows the normalized maximum ∂h=∂z nu-
occurring. The pressure profiles shown in Fig. 5(a) appear parallel, merically determined at the ground surface for all the S-LE cases
but they are not. The subtle differences can best be examined in (saturated condition, linear soil constitutive model). Notably, the

© ASCE 04019002-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


The B value can be selected manually without regard to the solid
and fluid bulk moduli, and different B values result in different
pore-water pressure head gradient profiles. Fig. 7 also shows the
reduced results for the numerical experiments investigating man-
ually selected B values. For the numerical experimentation per-
formed for manually selected B values, it was assumed that the
poroelasticity parameters are not affected by the effective stress
or porosity; that is, it was assumed that ν u ¼ 0.5 and α ¼ 1. As
the B value increases from zero to one, the amount of pore-water
pressure within the soil bed induced by the overlying tsunami water
weight increases, which results in lower values of ∂h=∂z. For the
limiting condition, when B ¼ 1, the values of ∂h=∂z become neg-
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

ligible; accordingly, when B ¼ 1, the overlying tsunami water


weight induces a change in excess pore-water pressure within
the soil bed equal to the weight of the overlying tsunami water,
which is the typical assumption employed when investigating
consolidation of soft clays (Terzaghi 1943).
Fig. 8 shows the variation of pore-water pressure and ∂h=∂z
with depth for the saturated soil condition at different times. To
create Fig. 8, the nonlinear soil constitutive model was used with
aired water as the pore fluid, i.e., the S-NI-UW Case from Table 2 is
investigated. Comparing Fig. 5 with Fig. 8 shows the difference
between the linear and nonlinear soil constitutive models in terms
of the pore-water pressure response in the soil bed due to the hypo-
thetical tsunami loading. At first glance, Figs. 5(a) and 8(a) look the
same and they are very similar; however, there are some differences
in the two figures best examined by investigating the differences in
∂h=∂z, which are shown in Figs. 5(b) and 8(b) for the linear and
nonlinear soil constitutive models, respectively. Primarily, the non-
linear soil constitutive model leads to values of ∂h=∂z that are non-
linear near the soil bed surface. The deformation near the soil bed
surface increases for a given overlying tsunami water weight as the
shear modulus decreases, which results in an increased value of
∂h=∂z near the soil bed surface, as observed in Fig. 8(b).
Fig. 9 shows the variation of ∂h=∂z with time at the ground
Fig. 4. Convergence in terms of excess pore-water pressure head gra- surface, 1 m below the ground surface, and 5 m below the ground
dient profiles for depths between 0 and 3 m for times of (a) 10 min; surface for Case S-NI-UW. Fig. 9 shows that ∂h=∂z reduces and
(b) 20 min; (c) 28 min; and (d) 29 min. becomes negative during the tsunami run-up, then ∂h=∂z increases
and becomes positive during tsunami drawdown, as expected based
on similar observations from Fig. 6. Fig. 9 also shows that ∂h=∂z
recorded at the ground surface and 1 m below the ground surface
Table 3. Summary of convergence studies with four different grids are relatively similar, although less similar than shown for Case
Grid Δz (m) mz Δt (s) N S-LE-UW (Fig. 6). In addition, the values of ∂h=∂z are larger near
the soil surface but smaller at a depth of 5 m during tsunami draw-
1 0.5 20 0.3193 5,900
2 0.1 100 0.0639 29,499
down when the nonlinear soil constitutive model is employed. The
3 0.05 200 0.0319 58,998 preceding observation highlights the effect of the shear modulus on
4 0.025 400 0.0160 117,996 the pore-water pressure response of a soil bed subjected to tsunami
loading. Careful measurements of the soil bed’s shear modulus
Note: mz = total number of cells; N = total number of time increments;
(and other elastic moduli) must be made to yield better predictions
Δt = time increment; and Δz = cell size.
of ∂h=∂z and thus, better predictions of sediment instability during
tsunami loading.
Fig. 10 shows the maximum values of ∂h=∂z for the nonlinear
maximum values of ∂h=∂z occur at the end of tsunami drawdown soil constitutive model cases in Table 2. Fig. 10 shows that the
for all the cases examined. Examining the bar heights for Case maximum value of ∂h=∂z is almost negligible for the case when
S-LE-PW and Case S-LE-UW in Fig. 7 shows the differences in deaired water is the pore fluid, compared with the case when aired
the response of soil beds saturated with deaired water (PW) and water is the pore fluid. More specifically, the maximum values of
aired water (UW). The difference between the two cases is signifi- ∂h=∂z at the ground surface during tsunami drawdown for the
cant; in fact, the maximum ∂h=∂z recorded at the ground surface cases when deaired water and aired water are the pore fluids are
when aired water is the pore fluid is much larger than the same 0.0032 and 0.1864, respectively. Fig. 10 also shows the maximum
measure when deaired water is the pore fluid. The difference shows values of ∂h=∂z for manually selected B values for the nonlinear
the importance of the bulk modulus of water. The B value depends soil constitutive model cases in Table 2. Similar to the linear soil
on the fluid bulk modulus, and because deaired water has a larger constitutive model cases, when the B value was selected manually,
bulk modulus than aired water (Verruijt 1969; Yang 2005; Young it was assumed that ν u ¼ 0.5 and α ¼ 1. In addition, similar to the
et al. 2009), it also has larger B values. linear soil constitutive model case, the maximum values of ∂h=∂z

© ASCE 04019002-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Case S-LE-UW: (a) total pore-water pressure at different times with depth for the saturated condition, assuming aired water and considering a
linear constitutive model; and (b) pressure head gradient at different times with depth for the saturated condition, assuming aired water and
considering a linear constitutive model.

Fig. 6. Case S-LE-UW: Pressure head gradient at different depths with


time for the saturated condition, aired water, and considering a linear
constitutive model.

increase as the B value decreases, which highlights the importance


of the overlying tsunami water weight.
Fig. 11 shows that the maximum values of ∂h=∂z at the ground
surface reduce linearly with reductions of the B value. As seen in
Fig. 11, the variation of the maximum ∂h=∂z at the ground surface
linearly changes with B value for both the linear and nonlinear soil Fig. 7. Maximum pressure head gradient at the ground surface for the
constitutive models. As the B value decreases, the nonlinear soil saturated condition for different cases using a linear constitutive model
constitutive model results in larger values of maximum ∂h=∂z and impermeable layer depth of 10 m below the ground’s surface.
at the ground surface compared with the linear soil constitutive The values are normalized with the maximum value, which occurs
model. In addition, assuming deaired water as the pore fluid results for Case S-LE-B(0) and is equal to 0.1628.
in negligible maximum ∂h=∂z at the ground surface for both soil
constitutive models. The relative compressibility between the pore
fluid and soil skeleton indicates how much excess pore-water pres- between soil skeleton and pore fluid. The B value in Fig. 11 was
sure is induced during tsunami loading (i.e., how the load is shared chosen manually; therefore, it is possible to manually control how
between the soil skeleton and how the pore water affects the pore- the load is shared between solid and fluid phases. Accordingly, in
water pressure response). The B value shows how the load is shared Fig. 11, the only difference between the nonlinear and linear cases

© ASCE 04019002-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Case S-NI-UW: (a) pore-water pressure changes with depth at different times for saturated nonlinear model and aired water; and (b) pressure
head gradient changes with depth at different times for saturated nonlinear model and aired water.

Fig. 9. Case S-NI-UW: Pressure head gradient changes with time at


different depths for saturated nonlinear model and aired water.

is the effect of the shear modulus. When the pore-water pressure


head gradient increases, the shear modulus reduces and the soil bed
deforms more when loaded by the tsunami water. Consequently,
when the shear modulus reduces, the magnitude of the pore-water
pressure head gradient increases.

Conclusions Fig. 10. Maximum pressure head gradient at the ground surface for the
saturated condition for different cases using a nonlinear constitutive
Tsunami loading causes partial drainage in soil beds. Moreover, model and impermeable layer depth of 10 m below the ground’s sur-
the loading is by a fluid, so seepage is important, and the loading face. The values are normalized with the maximum value, which occurs
timescale is rapid and transient. The authors presented a coupled for Case S-NI-B(0) and is equal to 0.2125.
seepage-deformation model considering the foregoing unique
tsunami characteristics. The model was implemented in MATLAB
using a finite-difference approach. The proposed model predicted
the pore-water pressure changes in soil beds due to idealized experiments focused on estimating the total pore-water pressure,
tsunami loading. pw , and excess pore-water pressure head gradient, ∂h=∂z. Different
Numerical experiments were performed on a beach sand to cases were examined. First, deaired water versus aired water as the
test the proposed seepage-deformation model. The numerical pore fluid was considered. Results showed that aired water caused

© ASCE 04019002-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


models to account for bed shear stresses and two-dimensional
seepage; (2) incorporating unsaturated soil conditions near the soil
surface (Abdollahi 2017); (3) considering earthquake–tsunami
multihazard (Abdollahi et al. 2017); (4) performing parametric
studies to understand the sensitivity of results to changes in input
parameters; and (5) consideration of the generation of excess pore-
water pressure during cyclic deformation under wave loading.
Finally, the hypothetical tsunami was chosen to represent a
realistic tsunami, particularly, the near-field tsunami anticipated
to attack the Pacific Northwest coast. None of the predicted values
of ∂h=∂z would cause liquefaction during the hypothetical
tsunami. Yet, enhanced scour and sediment transport is expected.
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

Moreover, the authors emphasize that the results are very sensitive
to the input soil conditions. Geotechnical and coastal engineers
should determine near-surface sediment properties to improve
the predictions. The work is formidable, but paramount for protect-
ing coastal areas during tsunami attack.

Appendix I. Deformation Model Modifications

Starting with the Fredlund et al. (2012) deformation model in


Fig. 11. Linear variation of the maximum pressure head gradient at the Eq. (1), the assumptions in Table 1 are used to reduce the functional
ground surface with B value for the saturated condition using linear and form for tsunami loading of soil beds. Additional assumptions not
nonlinear constitutive model. The impermeable layer depth is 10 m explicitly contained in Table 1 are stated within this Appendix.
below the ground’s surface. Considering that the soil bed is fully saturated and the strain
condition is uniaxial leads to pa ¼ 0 as well as the functional form
for H given in Appendix III. With the preceding assumptions,
Eq. (1) reduces to
larger ∂h=∂z values, and the B value changed with entrapped air
within the pore fluid. Second, the B values were differed, and re- δϵv ¼ mv ðδσ − δpw Þ ¼ mv δσ 0 ð11Þ
sults showed that the greatest values of ∂h=∂z increased as the B which explains the stress-strain behavior of one-dimensional con-
value decreased. Third, linear versus nonlinear soil stiffness models solidation testing (Lambe and Whitman 1969). Eq. (11) is modified
were tested. The nonlinear soil stiffness model led to larger values with the α term, following the Skempton (1984) effective stress
of ∂h=∂z near the soil surface. During momentary liquefaction, un- model shown in Appendix III, which leads to
like the residual liquefaction, reducing the B value increases the
liquefaction potential. Finally, during tsunami loading, the largest δϵv ¼ mv ðδσ − αδpw Þ ð12Þ
values of ∂h=∂z occurred near the ground surface, which indicates
that the effective stress reduction starts at the ground surface and For saturated conditions, the constitutive equations (Detournay
moves downward. The foregoing physical description is different and Cheng 1993) linking the total vertical stress, σ, to the vertical
from earthquake-induced liquefaction, which initiates at depths be- strain, ε, and total pore-water pressure, pw , (i.e., pore-water cou-
low the ground surface. pled constitutive equation) and linking the total vertical stress to the
The numerical experimentation showed trends with respect to volumetric water content, θw , (i.e., water content coupled constit-
pw and ∂h=∂z that appear physically reasonable, and the authors utive equation) are given, respectively, as follows:
are particularly encouraged by the proposed seepage-deformation
model’s simplicity and ability to incorporate partial drainage 2Gð1 − νÞ
σ¼ ϵ − αpw ð13aÞ
effects. Also, the model is applicable to tsunamis with different 1 − 2ν
sizes and tsunamis with more than one cycle. It is expected that
the model can be used in the future to predict tsunami-induced 2Gð1 − ν u Þ
σ¼ ϵ − αMδθw ð13bÞ
pore-water pressures in coastal sands, which can lead to better pre- 1 − 2ν u
dictions of tsunami-induced scour and sediment transport. Coastal
and geotechnical engineers can use the updated predictions, in turn, In Eq. (13b), δθw = increment of volumetric water content
to design and retrofit coastal buildings and infrastructure. In addi- (which shows the variation of fluid volume per unit volume of
tion, coastal geomorphologists can potentially use the model to up- porous media).
date calculations of sediment transport, which may lead to better Eqs. (13a) and (13b) are similar to the Biot (1941) poroelasticity
understanding of historical tsunami sizes. equations. To develop Eqs. (13a) and (13b), it was assumed that
Despite the promise that the proposed seepage-deformation the solid and fluid phases are fully connected, that the stress-strain
model holds, there are modeling limitations, which have largely relationship is linear and reversible, and that the porous medium is
been outlined in Table 1. Particularly, there is a paucity of high- isotropic. In addition, in Eqs. (13a) and (13b), the shear stress
quality field or large-scale laboratory data needed to validate the between the fluid phase and solid skeleton, as well as the fluid’s
model; therefore, in the near term, the authors recommend labora- viscosity, are neglected. Eqs. (13a) and (13b) are justified for quasi-
tory or field experiments to confirm the proposed seepage- static processes, as described by Detournay and Cheng (1993);
deformation model. Also, following the assumptions outlined in however, herein for the sake of simplicity, they are used it for rapid
Table 1, future work could focus on (1) furthering two-dimensional transient tsunami loading. When the undrained assumption is made,

© ASCE 04019002-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


the increment of volumetric water content, δθw , in Eq. (13b) goes If one considers a fully saturated soil bed governed by the linear
to zero. Because tsunami run-up and drawdown occur relatively soil stiffness model (i.e., G ¼ 35 MPa) with a porosity, φ, of 0.4
quickly, the undrained assumption is justified. Setting δθw equal to and using the model parameters shown in Appendix III and the
zero in Eq. (13b), then combining Eqs. (13a) and (13b) and solving foregoing arguments, then the dimensionless Zienkiewicz et al.
for pw yields (1980) Π parameters can be calculated as Π1 ≈ 170 and Π2 ≈
3.8 × 10−10 when deaired water is the pore fluid. Likewise, Π1 ≈
Bð1 þ ν u Þ
pw ¼ − σ ð14Þ 2.6 and Π2 ≈ 2.5 × 10−8 when aired water is the pore fluid
3ð1 − ν u Þ (i.e., 3% entrained air; K w ¼ 4 MPa). Accordingly, for the hypo-
Eq. (14) shows how much pore-water pressure is instantaneously thetical tsunami flow height time series and hypothetical soil bed
induced by the total stress due to weight of the tsunami. Subtracting considered herein, with the linear soil stiffness model, the dimen-
the pore-water pressure induced by the weight of the tsunami sionless Zienkiewicz et al. (1980) Π parameters indicate a slow
phenomenon, which indicates that the assumption to neglect the
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

[Eq. (14)], which removes the effect of the instantaneous hydrostatic


pore-water pressure (i.e., Assumption 4 in Table 1), results in inertial terms in the Biot (1941) equation is appropriate.
 
Bð1 þ ν u Þ
δϵv ¼ mv δσ − αδpw ð15Þ Validity of Darcy’s Law
3ð1 − ν u Þ
which is the final functional form as shown in Eq. (2). Eq. (3) was developed with the assumption that fluid flows through
the porous medium according to Darcy’s law (Bear 1988).
Although, fluid flow through porous media deviates from Darcy’s
Appendix II. Additional Description of Assumptions law for turbulent flow and high-velocity flow (e.g., Firoozabadi and
Katz 1979; Guppy et al. 1982; Martins et al. 1990; Zeng and Grigg
2006), porous flow often occurs at low velocities (Bear 1988).
According to Bear (1988), Darcy’s law is valid when the Reynolds
Neglecting Fluid Acceleration Terms
number [i.e., R ¼ ðρvp d50 Þ=μ, where vp is the seepage velocity
Zienkiewicz et al. (1980) defined two dimensionless parameters, Π1 and d50 is mean grain diameter] does not exceed values between
and Π2 , to determine whether the pore-water pressure response in 1 and 10. Tsunami-induced seepage is expected to have a modest
soil subjected to different loading is fast, moderate, or slow. The de- Reynolds number (i.e., ≈0.2 for the modeling parameters consid-
termination of fast, moderate, or slow pore-water pressure response, ering herein if d50 ¼ 0.2 mm and the maximum hydraulic conduc-
in turn, leads to the determination of the appropriateness of different tivity is ≈0.1 cm=s); however, the tsunami-induced seepage flow is
simplifying assumptions when solving the Biot (1941) equations. rapid transient (i.e., not steady flow). Mongan (1985) validated the
The dimensionless parameters, Π1 and Π2 , are given as follows: application of Darcy’s law to transient flow problems for most
 2 groundwater flow conditions; however, to the authors’ knowledge,
2k T T̂
Π1 ¼ d s2 ; Π2 ¼ π 2 ð16Þ the applicability of Darcy’s law to problems concerning tsunami-
πβgT̂ Ts induced seepage through porous media has not been explicitly
where β = density ratio (β ¼ ρ=ρb , where ρb is the soil bed’s mass examined.
density, i.e., ρb ¼ 1,990 kg=m3 ); kd = soil bed’s kinematic per- A more accurate method for calculating the total head of com-
meability; T s = tsunami wave period; and T̂ = natural period of the pressible fluids was given by Hubbert (1940). The Hubbert (1940)
soil bed, which Zienkiewicz et al. (1980) defined as T̂ ¼ ð2LÞ=V c , potential, Φ, is
where L is a length of interest and V c is the compressional wave Z
velocity of the fully saturated soil bed. The hypothetical tsunami
p dp
Φ ¼ gz þ ð19Þ
in Fig. 2(b) has a wave period T s of approximately 32 min, and p0 ρ
the length of interest L is the depth-to-bedrock, which is 10 m.
The compressional wave velocity V c depends on the soil bed’s where in this case, p = pressure applied by the tsunami water; and
and water’s bulk moduli and is given by Zienkiewicz et al. p0 = initial pressure. The water’s mass density, ρ, changes as a
(1980) as follows: function of the water’s bulk modulus. For a 15 m tsunami height
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi the pressure applied to the soil surface is approximately 150 kPa.
K þ Kϕw For the case of 3% entrained air (Appendix III), assuming that
Vc ¼ ð17Þ p0 ¼ 0 at the ground surface, then the second term on the right-
ρb
hand side of Eq. (19) is approximately 3 kPa. Accordingly, the
The kinematic permeability, kd , is given as follows (Bear 1988): gravity term in Eq. (19) dominates. For the foregoing reason and
to keep in line with this study’s objective of model simplicity, the
ρgkab validity of the typical calculation of total head (i.e., incompressible
kd ¼ ð18Þ
μ fluid assumption) when applying Darcy’s law is accepted.

Appendix III. Fundamentals Required for the Proposed Seepage-Deformation Model

Symbol Parameter Definition/value References


mv Coefficient of volume compressibility 1 − 2ν Lambe and Whitman (1969)
2Gð1 − νÞ
υ Drained Poisson’s ratio 0.3 Lambe and Whitman (1969)
G Shear modulusa h σ 0
n i Sawangsuriya et al. (2009) and Lu and Kaya (2014)
35 MPa; Aσr
σr

© ASCE 04019002-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


Appendix III. (Continued.)
Symbol Parameter Definition/value References
0
σ Effective stress σ − αpw Skempton (1984)
σr Reference stress 1 atm —
A Shear modulus fitting parameter 500 Sawangsuriya et al. (2009) and Lu and Kaya (2014)
n Shear modulus fitting parameter 0.5 Hardin and Richart (1963) and Janbu (1963)
K
α Effective stress modifierb 1− Skempton (1984)
Ks
2Gð1 þ νÞ
H Soil skeleton’s bulk modulus Biot (1941) and Lambe and Whitman (1969)
ð1 − 2νÞ
1 1
K − Ks
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

B B valuec Zienkiewicz and Bettess (1982),


1
K − 1þϕ ϕ
Ks þ Kw Kokusho (2000), and Yang (2005)
2Gð1 þ νÞ
K Soil bed’s bulk modulus Skempton (1984)
3ð1 − 2νÞ
Ks Soil grain’s bulk modulus 36 GPa Stoll and Kan (1981)
Kw Pore-water’s bulk modulusd (4 MPa, 2.2 GPa) Young et al. (2009)
3K u − 2G
υu Undrained Poisson’s ratio Detournay and Cheng (1993)
2ð3K u þ GÞ
Kð1 þ α2 K w Þ
Ku Undrained soil bed’s bulk modulus Detournay and Cheng (1993)
ð1 − αÞðα − ϕÞK w þ ϕK
ρ Water’s mass densitye 1,000 kg=m3 Granger (1985)
S Degree of saturation 1 —
kab Absolute permeabilityf 1 × 10−11 m2 Bear (1988)
kr Relative permeability 1 Bear (1988)
μ Water’s dynamic viscosity 1 × 10−3 Pa · s Granger (1985)
g Acceleration due to gravity 9.806 m=s2 —
a
Linear (i.e., G ¼ 35 MPa) and nonlinear stiffness models were used for the shear modulus. The nonlinear stiffness model is based on a profile measured by
Abdoun et al. (2013) for a uniform sand deposit. Table 1 and the “Experimental Plan” subsection offer more discussion.
b
Skempton (1984) reported α values of 0.9985 and 0.9997 for loose and dense sands, respectively.
c
A functional form for the B value is used that differs from the original functional form proposed by Skempton (1954) because it includes the effects caused by
the compressibility of the soil grains (Bishop 1973). The functional form can also be expressed as B ¼ ½3ðν u − νÞ=½αð1 − 2νÞð1 þ ν u Þ, as discussed by
Detournay and Cheng (1993).
d
K w depends on the entrained air within the water. Young et al. (2009) reported values of 4 MPa to 2.2 GPa for water with 3% entrained air and deaired water,
respectively.
e
A mass density of fresh water at approximately 4°C was assumed for the calculations herein instead of that for saltwater or sediment-laden saltwater; the
results are not very sensitive to the mass density.
f
The fact that kab is a constant value is a simplifying assumption, which is discussed in Table 1.

Acknowledgments Balhoff, M. T., S. G. Thomas, and M. F. Wheeler. 2008. “Mortar coupling


and upscaling of pore-scale models.” Comput. Geosci. 12 (1): 15–27.
The authors would like to acknowledge the input of Harry Yeh, https://doi.org/10.1007/s10596-007-9058-6.
Malgorzata Peszyńska, Solomon Yim, and T. Matthew Evans, who Bear, J. 1988. Dynamics of fluids in porous media. 1st ed. New York:
helped to strengthen the work. The authors were funded by the Dover.
National Science Foundation (NSF) under Grant No. CMMI- Bernard, E. N., and A. R. Robinson, eds. 2009. “Tsunamis.” Vol. 15 of
1538211, the Oregon State University College of Engineering, The sea. Cambridge, MA: Harvard University Press.
and the Cascadia Lifelines Program (CLiP). The financial support Biot, M. A. 1941. “General theory of three-dimensional consolidation.”
is acknowledged gratefully. J. Appl. Phys. 12 (2): 155–164. https://doi.org/10.1063/1.1712886.
Biot, M. A. 1956. “Theory of deformation of a porous viscoelastic aniso-
tropic solid.” J. Appl. Phys. 27 (5): 459–467. https://doi.org/10.1063/1
.1722402.
References Bishop, A. W. 1954. “The use of pore-pressure coefficients in practice.”
Abdollahi, A. 2017. “Tsunami-induced pore water pressure changes in Geotechnique 4 (4): 148–152. https://doi.org/10.1680/geot.1954.4.4
soil beds: Numerical formulation and experimentation.” Ph.D. thesis, .148.
School of Civil and Construction Engineering, Oregon State Univ. Bishop, A. W. 1973. “The influence of an undrained change in stress on the
Abdollahi, A., R. K. Adams, and H. B. Mason. 2017. “Earthquake-induced pore pressure in porous media of low compressibility.” Geotechnique
water level rise: Implications for tsunami-induced sediment instability 23 (3): 435–442. https://doi.org/10.1680/geot.1973.23.3.435.
in coastal areas.” In Proc., Geotechnical Frontiers 2017, 428–438. Bukač, M., I. Yotov, R. Zakerzadeh, and P. Zunino. 2015. “Partitioning
Reston, VA: ASCE. strategies for the interaction of a fluid with a poroelastic material based
Abdoun, T., M. Gonzalez, S. Thevanayagam, R. Dobry, A. Elgamal, M. on a Nitsches coupling approach.” Comput. Methods Appl. Mech. Eng.
Zeghal, V. Mercado, and U. El Shamy. 2013. “Centrifuge and 292: 138–170. https://doi.org/10.1016/j.cma.2014.10.047.
large-scale modeling of seismic pore pressures in sands: Cyclic strain Carrier, G. F., T. T. Wu, and H. Yeh. 2003. “Tsunami run-up and draw-
interpretation.” J. Geotech. Geoenviron. Eng. 139 (8): 1215–1234. down on a plane beach.” J. Fluid Mech. 475: 79–99. https://doi.org/10
https://doi.org/10.1061/(ASCE)GT.1943-5606.0000821. .1017/S0022112002002653.

© ASCE 04019002-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


Chen, H., and X.-P. Wang. 2014. “A one-domain approach for modeling Liu, X., M. H. García, and R. Muscari. 2007. “Numerical investigation
and simulation of free fluid over a porous medium.” J. Comput. Phys. of seabed response under waves with free-surface water flow.” Int. J.
259: 650–671. https://doi.org/10.1016/j.jcp.2013.12.008. Offshore Polar Eng. 17 (2): 97–104.
Chowdhury, B., G. Dasari, and T. Nogami. 2006. “Laboratory study of Lowe, D. R. 1975. “Water escape structures in coarse-grained sediments.”
liquefaction due to wave-seabed interaction.” J. Geotech. Geoenviron. Sedimentology 22 (2): 157–204. https://doi.org/10.1111/j.1365-3091
Eng. 132 (7): 842–851. https://doi.org/10.1061/(ASCE)1090-0241 .1975.tb00290.x.
(2006)132:7(842). Lowe, J., and L. Karafiath. 1960. “Effect of anisotropic consolidation on
Detournay, E., and A. H.-D. Cheng. 1993. “Fundamentals of poroelastic- the undrained shear strength of compacted clays.” In Proc., ASCE Re-
ity.” In Vol. 2 of Comprehensive rock engineering: Principles, practice search Conf. on Shear Strength of Cohesive Soils, 837–858. Reston,
and projects, edited by J. A. Hudson, 113–171. Oxford, UK: Pergamon. VA: ASCE.
Discacciati, M., E. Miglio, and A. Quarteroni. 2002. “Mathematical and Lu, N., and M. Kaya. 2014. “Power law for elastic moduli of unsaturated
numerical models for coupling surface and groundwater flows.” Appl. soil.” J. Geotech. Geoenviron. Eng. 140 (1): 46–56. https://doi.org/10
Numer. Math. 43 (1–2): 57–74. https://doi.org/10.1016/S0168-9274 .1061/(ASCE)GT.1943-5606.0000990.
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

(02)00125-3. Martins, J. P., D. Milton-Tayler, and H. K. Leung. 1990. “The effects of


Duncan, J. M., S. Wright, and K. S. Wong. 1990. “Slope stability during non-Darcy flow in propped hydraulic fractures.” In Proc., Society
rapid drawdown.” In Vol. 2 of Proc., H. Bolton Seed Memorial Symp., of Petroleum Engineers Annual Technical Conf. and Exhibition.
edited by J. M. Duncan, 253–272. Vancouver, BC, Canada: Bitech. Richardson, TX: Society of Petroleum Engineers.
Firoozabadi, A., and D. L. Katz. 1979. “An analysis of high-velocity gas Mason, H. B., and H. Yeh. 2016. “Sediment liquefaction: A pore-water
flow through porous media.” J. Pet. Technol. 31 (2): 211–216. https:// pressure gradient view-point.” Bull. Seismol. Soc. Am. 106 (4):
doi.org/10.2118/6827-PA. 1908–1913. https://doi.org/10.1785/0120150296.
Fredlund, D. G., and N. Morgenstern. 1976. “Constitutive relations for Mayr, M., T. Kloppel, W. A. Wall, and M. W. Gee. 2015. “A temporal
volume change in unsaturated soils.” Can. Geotech. J. 13 (3): consistent monolithic approach to fluid-structure interaction enabling
261–276. https://doi.org/10.1139/t76-029. single field predictors.” SIAM J. Sci. Comput. 37 (1): B30–B59.
Fredlund, D. G., H. Rahardjo, and M. D. Fredlund. 2012. Unsaturated soil https://doi.org/10.1137/140953253.
mechanics in engineering practice. 1st ed. Hoboken, NJ: Wiley. Mei, C. C., and M. A. Foda. 1981. “Wave-induced responses in a fluid-
Fredlund, M., H. Lu, and T. Feng. 2011. “Combined seepage and slope filled poro-elastic solid with a free surface: A boundary layer theory.”
stability analysis of rapid drawdown scenarios for levee design.” Geophys. J. Int. 66 (3): 597–631. https://doi.org/10.1111/j.1365-246X
In Proc., Geo-Frontiers 2011: Advances in Geotechnical Engineering, .1981.tb04892.x.
1604–1595. Reston, VA: ASCE. Merxhani, A., and D. Liang. 2012. “Investigation of the poro-elastic
Fritz, H. M., D. A. Phillips, A. Okayasu, T. Shimozono, H. Liu, F. response of seabed to solitary waves.” In Proc., 22nd Int. Offshore
Mohammed, V. Skanavis, C. E. Synolakis, and T. Takahashi. 2012. and Polar Engineering Conf. Mountain View, CA: International
“The 2011 Japan tsunami current velocity measurements from survivor Society of Offshore and Polar Engineers.
videos at Kesennuma Bay using lidar.” Geophys. Res. Lett. 39 (7): Mongan, C. E. 1985. Validity of Darcy’s law under transient conditions.
L00G23. https://doi.org/10.1029/2011GL050686. USGS Professional Paper 1331. Washington, DC: United States
Goyeau, B., D. Lhuillier, D. Gobin, and M. Velarde. 2003. “Momentum Geological Survey.
transport at a fluid: Porous interface.” Int. J. Heat Mass Transfer Mosthaf, K., K. Baber, B. Flemisch, R. Helmig, A. Leijnse, I. Rybak, and
46 (21): 4071–4081. https://doi.org/10.1016/S0017-9310(03)00241-2. B. Wohlmuth. 2011. “A coupling concept for two-phase compositional
Granger, R. A. 1985. Fluid mechanics. 1st ed. New York: Dover. porous-medium and single-phase compositional free flow.” Water
Gratiot, N., and M. Mory. 2000. “Wave-induced sea bed liquefaction with Resour. Res. 47 (10): 1–19. https://doi.org/10.1029/2011WR010685.
application to mine burial.” In Proc., 10th Int. Offshore and Polar En- Ochoa-Tapia, J. A., and S. Whitaker. 1995. “Momentum transfer at the
gineering Conf. Mountain View, CA: International Society of Offshore boundary between a porous medium and a homogeneous fluid: Theo-
and Polar Engineers. retical development.” Int. J. Heat Mass Transfer 38 (14): 2635–2646.
Guppy, K. H., H. Cinco-Ley, and H. J. Ramey. 1982. “Pressure buildup https://doi.org/10.1016/0017-9310(94)00346-W.
analysis of fractured wells producing at high flow rates.” J. Pet. Peaceman, D. W. 1977. Fundamentals of numerical reservoir simulation.
Technol. 34 (11): 2656–2666. https://doi.org/10.2118/10178-PA. 1st ed. New York: Elsevier.
Hardin, B. O., and F. E. J. Richart. 1963. “Elastic wave velocities in granu- Peszyńska, M., E. W. Jenkins, and M. F. Wheeler. 2002. “Boundary
lar soils.” J. Soil Mech. Found. Div. 89 (1): 33–65. conditions for fully implicit two-phase flow.” In Vol. 306 of Recent ad-
Helmig, R. 1997. Multiphase flow and transport processes in the subsur- vances in numerical methods for partial differential equations and ap-
face: A contribution to the modeling of hydrosystems. 1st ed. Berlin: plications, edited by X. Feng and T. P. Schulze, 85–106. Providence, RI:
Springer. American Mathematical Society.
Hubbert, M. K. 1940. “The theory of ground-water motion.” J. Geol. 48 (8): Peszyńska, M., and S.-Y. Yi. 2008. “Numerical methods for unsaturated
785–944. https://doi.org/10.1086/624930. flow with dynamic capillary pressure in heterogeneous porous media.”
Janbu, N. 1963. “Soil compressibility as determined by oedometer Int. J. Numer. Anal. Model. 5: 126–149.
and triaxial tests.” In Vol. 1 of Proc., 2nd European Conf. on Soil Putnam, J. 1949. “Loss of wave energy due to percolation in a permeable
Mechanics and Foundation Engineering, 19–25. Wiesbaden, Germany: sea bottom.” EOS Trans. Am. Geophys. Union 30 (3): 349–356. https://
Soil Mechanics Foundation. doi.org/10.1029/TR030i003p00349.
Kirca, V. O., B. M. Sumer, and J. Fredsøe. 2013. “Residual liquefaction of Richards, L. A. 1931. “Capillary conduction of liquids through porous
seabed understanding waves.” J. Waterw. Port Coastal Ocean Eng. mediums.” J. Appl. Phys. 1 (5): 318–333.
139 (6): 489–501. https://doi.org/10.1061/(ASCE)WW.1943-5460 Sakai, T., A. Hattori, and K. Hatanaka. 1991. “Wave-induced transient
.0000208. pore-water pressure and seabed instability in the surf zone.” In Proc.,
Kokusho, T. 2000. “Correlation of pore-pressure B-value with P-wave Int. Conf. on Geotechnical Engineering for Coastal Development-
velocity and Poisson’s ratio for imperfectly saturated sand or gravel.” Theory and Practice on Soft Ground, 627–632. Tokyo: Coastal
Soils Found. 40 (4): 95–102. https://doi.org/10.3208/sandf.40.4_95. Development Institute of Technology.
Koshimura, S., and S. Hayashi. 2012. “Tsunami flow measurement using Santamarina, J. C., K. A. Klein, and M. A. Fam. 2001. Soils and waves. 1st
the video recorded during the 2011 Tohoku tsunami attack.” In Proc., ed. Chichester, UK: Wiley.
IEEE Int. Geoscience and Remote Sensing Symp., 6693–6696. Sawangsuriya, A., T. B. Edil, and P. J. Bosscher. 2009. “Modulus-suction-
New York: IEEE. moisture relationship for compacted soils in postcompaction state.”
Lambe, T. W., and R. V. Whitman. 1969. Soil mechanics. 1st ed. New York: J. Geotech. Geoenviron. Eng. 135 (10): 1390–1403. https://doi.org/10
Wiley. .1061/(ASCE)GT.1943-5606.0000108.

© ASCE 04019002-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002


Seed, H. B. 1979. “Soil liquefaction and cyclic mobility evaluation for level Yeh, H., and S. Sato. 2016. “Tsunami effects on buildings and coastal struc-
ground during earthquakes.” J. Geotech. Eng. Div. 105 (2): 201–255. tures (special issue on uncertainties in tsunami effects).” J. Disaster Res.
Skempton, A. W. 1954. “The pore-pressure coefficients A and B.” Geotech- 11 (4): 662–669. https://doi.org/10.20965/jdr.2016.p0662.
nique 4 (4): 143–147. https://doi.org/10.1680/geot.1954.4.4.143. Young, Y. L., J. A. White, H. Xiao, and R. I. Borja. 2009. “Liquefaction
Skempton, A. W. 1984. “Effective stress in soils, concrete and rocks.” potential of coastal slopes induced by solitary waves.” Acta Geotechn-
In Selected papers on soil mechanics by A. W. Skempton, edited by ica 4 (1): 17–34. https://doi.org/10.1007/s11440-009-0083-6.
J. B. Burland, 106–118. London: Thomas Telford. Zarba, R. L., E. Bouloutas, and M. Celia. 1990. “General mass-
Stoll, R. D., and T.-K. Kan. 1981. “Reflection of acoustic waves at a water- conservative numerical solution for the unsaturated flow equation.”
sediment interface.” J. Acoust. Soc. Am. 70 (1): 149–156. https://doi.org Water Resour. Res. 26 (7): 1483–1496. https://doi.org/10.1029
/10.1121/1.386692. /WR026i007p01483.
Sumer, B. M. 2014. Liquefaction around marine structures. 1st ed. Zen, K., and H. Yamazaki. 1990a. “Mechanism of wave-induced liquefac-
Singapore: World Scientific. tion and densification in seabed.” J. Soils Found. 30 (4): 90–104. https://
Sumer, B. M., M. B. Sen, I. Karagali, B. Ceren, J. Fredsøe, M. Sottile, L.
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 04/10/19. Copyright ASCE. For personal use only; all rights reserved.

doi.org/10.3208/sandf1972.30.4_90.
Zilioli, and D. R. Fuhrman. 2011. “Flow and sediment transport induced
Zen, K., and H. Yamazaki. 1990b. “Oscillatory pore pressure and liquefac-
by a plunging solitary wave.” J. Geophys. Res.: Oceans 116 (C1): 1–15.
tion in seabed induced by ocean waves.” J. Soils Found. 30 (4):
https://doi.org/10.1029/2010JC006435.
147–161. https://doi.org/10.3208/sandf1972.30.4_147.
Terzaghi, K. 1943. Theoretical soil mechanics. 1st ed. New York: Wiley.
Verruijt, A. 1969. “Elastic storage of aquifers.” In Flow through porous Zeng, Z., and R. Grigg. 2006. “A criterion for non-Darcy flow in porous
media. 1st ed., edited by R. J. M. DeWiest, 331–376. New York: media.” Transp. Porous Media 63 (1): 57–69. https://doi.org/10.1007
Academic Press. /s11242-005-2720-3.
Woodward, C. S., and C. N. Dawson. 2000. “Analysis of expanded mixed Zienkiewicz, O. C., and P. Bettess. 1982. “Soils and other saturated media
finite element methods for a nonlinear parabolic equation modeling under transient, dynamic conditions; general formulation and the
flow into variably saturated porous media.” SIAM J. Numer. Anal. validity of various simplifying assumptions.” In Soil mechanics: Tran-
37 (3): 701–724. https://doi.org/10.1137/S0036142996311040. sient and cyclic loads, edited by O. C. Zienkiewicz and G. N. Pande,
Yang, J. 2005. “Pore pressure coefficient for soil and rock and its relation to 1–16. Chichester, UK: Wiley.
compressional wave velocity.” Geotechnique 55 (3): 251–256. https:// Zienkiewicz, O. C., A. Chan, M. Pastor, B. Schrefler, and T. Shiomi. 1999.
doi.org/10.1680/geot.2005.55.3.251. Computational geomechanics. Chichester, UK: Wiley.
Yeh, H., and H. B. Mason. 2014. “Sediment response to tsunami loading: Zienkiewicz, O. C., C. Chang, and P. Bettess. 1980. “Drained, undrained,
Mechanisms and estimates.” Geotechnique 64 (2): 131–143. https://doi consolidating and dynamic behaviour assumptions in soils.” Geotech-
.org/10.1680/geot.13.P.033. nique 30 (4): 385–395. https://doi.org/10.1680/geot.1980.30.4.385.

© ASCE 04019002-12 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2019, 145(3): 04019002

You might also like