Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

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

net/publication/325968284

Soil-Water-Structure Interaction Algorithm in Smoothed Particle


Hydrodynamics (SPH) with Application to Deep-Penetrating Problems

Article in International Journal of Computational Methods · June 2018


DOI: 10.1142/S0219876218501359

CITATIONS READS

7 1,599

4 authors, including:

Hao Wu
Shanghai Jiao Tong University
6 PUBLICATIONS 62 CITATIONS

SEE PROFILE

All content following this page was uploaded by Hao Wu on 07 October 2018.

The user has requested enhancement of the downloaded file.


2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

International Journal of Computational Methods


Vol. 15, No. 7 (2018) 1850135 (24 pages)
c World Scientific Publishing Company
DOI: 10.1142/S0219876218501359

Soil–Water-Structure Interaction Algorithm


in Smoothed Particle Hydrodynamics (SPH)
with Application to Deep-Penetrating Problems

Hao Wu∗ , Jian Wang†,§ , Chun Wang‡ and Jian Hua Wang∗
∗Department of Civil Engineering

Shanghai Jiao Tong University, P. R. China


†Department of Water Conservancy and

Hydropower Engineering, Hohai University, P. R. China


‡School of Naval Architecture, Ocean and Civil Engineering
Shanghai Jiao Tong University, P. R. China
§chulizi hhu@hotmail.com

Received 24 March 2018


Revised 3 June 2018
Accepted 18 June 2018
Published

A fully coupled soil–water-structure interaction algorithm was presented in the frame-


work of smoothed particle hydrodynamics (SPH). In this algorithm, soil–water interac-
tion was simulated based on the two-phase mixture theory. Each phase of the mixture
occupies part of the macroscopic mixture and satisfies its own conservation equations of
mass and momentum. The Drucker–Prager model with nonassociated plastic flow rule
was used to describe the constitutive behavior of soil. The water was treated as Newto-
nian fluid. Interaction between soil and water was modeled by the pore water pressure
and the viscous drag force. The structure was considered as rigid and the interaction with
soil/water was modeled by the frictional sliding contact algorithm. With this algorithm,
it is possible to investigate pore water pressure, the effective stress and deformation of
the soil undergoing large deformation. Moreover, the effect of the temporal and spatial
evolution of soil porosity was taken into consideration. This study first examined the
proposed algorithm for a U-tube seepage problem and a two-dimensional consolidation
problem. Afterwards, the continuous deep penetrating process of the spudcan, which
involved large soil deformation and complex soil–water-structure interaction, was sim-
ulated under axisymmetric conditions The comparison with previous research indicates
the robustness and applicability of the proposed algorithm. Furthermore, the proposed
approach could be a potentially efficient tool helping to reveal the mechanism of soil
failure in geotechnical, costal and ocean engineering.

Keywords: Soil–water-structure interaction; SPH; frictional sliding contact; deep


penetrating.

§ Corresponding author.

1850135-1
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

1. Introduction
Soil–water-structure interaction is a commonly seen and challenging problem
in geotechnical, coastal and ocean engineering. Typical examples include deep-
penetrating, scour and erosion, underwater excavation and liquefaction problems.
One of the most formidable challenges is to depict the sophisticated interac-
tion of soil with water. Single phase [Bagnold (1954)] and Quasi-single phase
[Iverson and Denlinger (2001)] treatments are restricted to simplified analysis as
a result of deviating from the real situation. The well-known Biot–Zienkiewicz
[Biot (1941); Zienkiewicz and Shiomi (1984)] consolidation theory takes water into
account implicitly in terms of pore water pressure. It has been widely applied to
consolidation and seepage problems, which focus on soil deformation or the station-
ary state. Less attention was paid to the flow process. Furthermore, the interaction
between saturated soil and water was ignored. Consequently, the Biot–Zienkiewicz
theory seems to be not suitable to deal with the aforementioned problems. In the
two-phase mixture theory [Drew (1983); Wang and Hutter (1999)], each constituent
occupies a certain amount of the volume space in the macroscopic mixture and sat-
isfies its own conservation equations of mass and momentum. One of the advantages
of the mixture theory over Biot–Zienkiewicz model is its ability to investigate the
pure fluid region and the interaction between pure fluid and saturated soil. Detailed
theoretical comparison between the mixture theory and Biot–Zienkiewicz theory
could be found in Coussy et al. [1998].
Another challenge of solving the problems mentioned above is to seek an appro-
priate numerical solution scheme. It has been proved that numerical results pre-
dicted by the conventional small strain analysis are significantly different from
Large Deformation Finite Element (LDFE) analysis when soil undergoes extremely
large deformation [Hossain and Randolph (2010)]. Although Arbitrary Lagrangian–
Eulerian (ALE) method and Coupled Eulerian–Lagrangian (CEL) method have
been utilized, numerical errors may still occur when encountering severe mesh dis-
tortion. The reason is mainly the very small time step resulted from the small
Lagrangian elements or Euler cells. Besides, determining the precise free sur-
face within the frame of fixed Euler grid is also a formidable work. Recently,
Smoothed Particle Hydrodynamics (SPH) has been extensively applied to astro-
physics [Gingold (1977)], fluid dynamics [Takeda et al. (1994); Liu and Li (2016)]
and computational geomechanics [Bui et al. (2008)] due to its mesh free nature.
Bui et al. [2007] and Bui and Fukagawa [2013] innovated the numerical simulation
of soil–water interaction using SPH method. They assumed that the saturated soil
could be divided into two separate phases. Nevertheless, soil and water particles
are not allowed to merge with each other in their model. The porosity in the
conservation equations was missing. Later, an improved mixture model that takes
porosity into account was developed and applied to seepage and liquefaction prob-
lems [Huang et al. (2013)]. Yet, the porosity was considered as a constant, which
may vary significantly in reality. As a matter of fact, the dynamics of the mixture

1850135-2
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

could be enormously influenced by the porosity. The case in point is the drag force
related to the porosity. To our knowledge, the only SPH mixture model considering
the effect of the spatially variable porosity was proposed by Wang et al. [2016a],
in which the volume fraction of water (essentially the porosity) was included in
the governing equations. Regrettably, the physical meaning of the so-called “stress
tensor of dry soil” was somewhat obscure. Relating it directly to the strain rate
in the constitutive model for soil phase appears to be inappropriate. In this study,
the total stress of the mixture will be decomposed according to Terzaghi’s effective
stress theory. The effective stress was explicitly utilized in the momentum equation
and obtained directly from the constitutive equation. The spatially variable poros-
ity was treated as a field variable together with other fundamental variables, i.e.,
density, velocity and stress.
In addition to the former two troublesome challenges, a robust contact algo-
rithm to model the interaction of the structure with mixture is required. In SPH,
field variables on a particle are interpolated over its neighbors within the support
domain. When the particle is located on or near the boundary, the support domain
is truncated by the boundary and not complete. The accuracy is then affected enor-
mously, which is the so-called boundary deficiency [Wang et al. (2013)]. The existent
boundary treatment will be reviewed later. The frictional contact algorithm that
origins in standard Discrete Element Method (DEM) will be adopted to determine
the boundary interaction force.
In the following sections, the mathematical formulation of the mixture theory
was first presented, followed by the constitutive model for water and soil. After
that, the basic SPH formulation and the governing equations discretized in SPH
form were proposed. Then the frictional sliding contact algorithm was introduced.
Afterwards, time integration and a few numerical issues including the treatment
of tensile ability were presented. Finally, typical numerical tests were conducted to
validate the proposed algorithm.

2. Governing Equations
Let’s consider the saturated soil–water mixture. The two-phase mixture theory
[Drew (1983); Wang and Hutter (1999)] assumes that each phase is present at any
point in the considered domain with different volume fractions. The volume frac-
tions are ϕs , ϕf are defined as the percentages of volume occupied by soil and water,
respectively. Possessing the identical physical meaning, the volume fraction of water
is replaced by the porosity n and (1 − n) is used instead of the volume fraction of
soil based on the saturation relationship.
The mass conservation equations are given in the following Lagrangian forms
[Wang et al. (2016a)]:

dρs
= −ρs ∇ · v s , (1)
dt

1850135-3
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

dρf
= −ρf ∇ · v f , (2)
dt
where ρs and ρf are the apparent density of soil and water, given by ρs = (1 − n)
ρ̃s , ρf = nρ̃f , respectively. ρ̃s is the particle density of soil, not including the pore
spaces between the grains, ρ̃f is the intrinsic density of the pore water. v s and v f
are the spatially averaged velocity of soil and water, respectively. By substituting
ρs with (1 − n)ρ̃s in Eq. (1) and considering the particle density ρ̃s as constant, the
time derivative of the porosity is given by
dn
= (1 − n)∇ · v s . (3)
dt
Now, let us decompose the total stress tensor σ of the mixture. In Wang et al.
[2016a], for soil, the partial stress tensor was defined as σ s = (1 − n)σ̃ s , where
σ̃ s is assumed to be the ‘stress tensor of dry soil ’. For water, the relationship
σ̃ f = −p̃f I+nτ̃ f is assumed, where p̃f and τ̃ f are the pore water pressure and shear
stress of water, respectively. The so-called ‘stress tensor of dry soil ’ σ̃s was assumed
to be directly related to the strain rate in the constitutive model for soil. It should
be noted that the physical meaning of σ̃ is somewhat nebulous and inconsistent
with the effective stress theory commonly used in geotechnical field. Here, we apply
the Terzaghi’s effective stress theory [Terzaghi (1943)]. The total stress tensor of
the mixture is resolved into the following form as:
σ = σ  − (1 − n)p̃f I + nσ̃ f , (4)
σ̃ f = −p̃f I + τ̃ f , (5)
where σ  is the effective stress used in the constitutive equation for soil. With these
assumptions, the conservation equations for momentum in Lagrangian form become
dv s
ρs = ∇ · σ  − (1 − n)∇p̃f + f d + ρs g, (6)
dt
dv f
ρf = −n∇p̃f + ∇ · (nτ̃ f ) − f d + ρf g. (7)
dt
By applying −n∇p̃f = −∇(np̃f ) + p̃f ∇n to Eq. (7), the conservation equation of
momentum for water can be rewritten as
dv f
ρf = −∇(np̃f ) + p̃f ∇n + ∇ · (nτ̃ f ) − f d + ρf g, (8)
dt
where g is the gravity acceleration, f d is the viscous drag force calculated through
the Darcy’s law. Assuming to be linear with respect to the velocity difference
between two phases, f d is obtained by Eq. (9),
n2 ρ̃f g(v f − v s )
fd = , (9)
k
where g is the magnitude of the gravity acceleration, k is the hydraulic conductivity
with the same dimensions as the velocity [LT−1 ].

1850135-4
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

In order to solve the above governing equations, constitutive model for both
water and soil are needed to determine the stress tensor. In this study, water is
assumed to be weakly compressible Newtonian fluid. The deviatoric stress of water
is given by Eq. (10),

τfαβ = µεαβ
f , (10)

where µ is the dynamic viscosity of water and εαβ


f is the shear strain rate, defined as

∂vfα ∂vfβ 2
εαβ
f = β
+ α
− (∇ · v f )δ αβ , (11)
∂x ∂x 3
where α and β denote the Cartesian components x, y, or z. δ αβ is the Kronecker
delta symbol. Note that the Einstein notation is used here. The pore water pressure
is calculated through density variation, i.e., equation of state, as follows:
 χ 
ρ̃f
p̃f = B −1 , (12)
ρ̃f 0
where B is a problem-dependent parameter that sets a limit to the maximum density
variation, χ is a constant normally set to 7 for water, ρ̃f 0 is the reference intrinsic
density of water.
In the current study, an elastic-perfectly plastic model is employed to describe
the soil behavior. Bui et al. pioneered the calculation of isotropic hydrostatic pres-
sure of soil directly from the constitutive model, whereas previous research adopted
the equation of state. For completeness, the final form of the constitutive model is
given in Eq. (13), more details could be found in [Bui et al. (2007)]. The Drucker–
Prager (D–P) yield criterion is used and a nonassociated flow rule is employed to
compute the plastic strain rate. Besides, the Jaumann stress rate is adopted to
consider a large deformation problem. To simplify the representation, the authors
make the convention that in the following sections drained parameters for soil will
be used without special identification. For instance, ψ is used instead of ψ  .

σ̇ αβ = σ αγ ω̇sβγ + σ γβ ω̇sαγ + 2Gėαβ


s
 
G αβ
+ K ε̇γγ
s δ
αβ
− λ̇ 9K sin ψδ αβ
+  s , (13)
J2

where the first two terms are the Jaumann stress rate, the next two terms are the
elastic part and the final one is the plastic part, ω̇s and ε̇s represent the spin and
strain rate tensor, given by Eqs. (14) and (15), respectively.
 
1 ∂vsα ∂vsβ
ω̇sαβ = − , (14)
2 ∂xβ ∂xα
 
αβ 1 ∂vsα ∂vsβ
ε̇s = + α , (15)
2 ∂xβ ∂x

1850135-5
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

where G is the shear modulus and K is the bulk modulus that relates to the Young’s
modulus E and the Poisson’s ratio υ, given by Eqs. (16) and (17)
E
G= , (16)
2(1 + υ)
E
K= , (17)
3(1 − 2υ)
where λ̇ is the rate of change of plastic multiplier, calculated through Eq. (18) as
  αβ αβ
3αφ K ε̇γγ
s + (G/ J2 )s ε̇s
λ̇ = , (18)
27αφ K sin ψ + G
where ψ is the dilatancy angle, αφ is the Drucker–Prager’s constants defined by
Eq. (19)
tan φ
αφ =  . (19)
9 + 12 tan2 φ
The yield criterion is expressed as Eq. (20).

f (I1 , J2 ) = J2 + αφ I1 − kc = 0, (20)
where I1 and J2 are the first and second stress invariants, given by I1 = σ γγ /3
and J2 = sαβ sαβ /2, respectively, sαβ is the deviatoric stress tensor of soil,
defined by sαβ = σ αβ − 13 σ γγ δ αβ , kc is the Drucker–Prager’s constants, calculated
by kc = √ 3c 2 , c and φ are the cohesion and internal friction angle of soil,
9+12 tan φ
respectively.

3. SPH Formulations
In the SPH method, by using a finite number of particles that carry field variables
and material properties, the computational domain is discretized and the govern-
ing equations are solved [Liu et al. (2008); Liu and Liu (2010)]. All the functional
variables are calculated through integral interpolation as

f (x) = f (x )W (x − x , h)dx , (21)

where W is the kernel or smoothing function, h is the smoothing length defining
the influence domain of W . The kernel W should be chosen as an even function and
satisfies the so-called normalizing condition, the delta condition, and the compact
condition. The Wendland kernel function is employed here for its accuracy and
efficiency, expressed as
(1 − q/2)4 (2q + 1), 0 ≤ q ≤ 2,
W (q, h) = αd × (22)
0, q ≥ 2,
where αd is the normalizing factor, αd = 7/4πh2 for two-dimension problems, q is
relative distance, q = |x − x |/h. The spatial derivatives of f (x) could be obtained

1850135-6
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

simply by substituting f (x) with ∇ · f (x). By using the divergence theorem, one
obtains

∇ · f (x) = f (x )W (x − x , h) · ndS


S

− f (x ) · ∇W (x − x , h)dx . (23)

With regard to particles associated with complete support domain, the surface
integral on the right-hand side of Eq. (23) is zero. Therefore, Eq. (23) is simplified as

∇ · f (x) = − f (x ) · ∇W (x − x , h)dx . (24)

Since the whole system in SPH is represented by a finite number of particles, which
not only act as interpolation points but also carry material properties. The integral
representations (expressed by Eqs. (21) and (24)) could be converted to the following
particle approximation form:
N
mj
f (xi ) = f (xj )W (xi − xj , h), (25)
j=1
ρj

N
mj
∇ · f (xi ) = − f (xj ) · ∇j W (xi − xj , h), (26)
j=1
ρj

where i and j denote particles, N is the total number of neighbor particles, m is


mass and ρ is the density, mj /ρj actually gives the finite volume ∆Vj that orig-
inates in the infinitesimal volume dx . It is worth noting that Eq. (26) can only
be applied to particles with complete support domain. As for particles on or near
the boundary, the support domain is truncated by the boundary and Eq. (26) is
no longer accurate. In the early application of SPH such as astrophysics and cos-
mology, boundary deficiency is usually either not important or requires very simple
treatment. However, in the mixture flow problems, both soil and water particles
interacting with the structure have to be treated properly. This will be discussed in
the next section.
In order to reduce errors arising from the particle inconsistency, several alterna-
tive SPH formulations for the spatial derivative of functions are derived as
N
1
∇ · f (xi ) = mj [f (xj ) − f (xi )] · ∇i W (xi − xj , h), (27)
ρi j=1

N
 
f (xi ) f (xj )
∇ · f (xi ) = ρi mj + · ∇i W (xi − xj , h). (28)
j=1
ρ2i ρ2j

By applying the SPH particle approximation to the gradients, the conservation


equations of mass and momentum can be rewritten in SPH form. Soil and water
particles are initially superimposed and then move separately according to their

1850135-7
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

own governing equations. Since there exist two different particles in the present
mixture model, i.e., soil and water particles, subscripts i and j are used to denote
soil variables whereas a and b denote the water variables.
The SPH formulation for porosity evaluation at a given soil particle i could be
rewritten as
N
dni mj
= (1 − ni ) v ji · ∇i Wij , (29)
dt ρ
j=1 j
N mj
where v ji = v j − v i . Making use of j=1 ρj · ∇i Wij = 0, the velocity difference
is included here to reduce errors arising from the particle inconsistency problem.
Calculation of density for soil particles is straightforward. As the particle density
ρ̃s of soil keeps unchanged, the apparent density of soil can be obtained through
ρs = (1 − n)ρ̃s once ni has been found.
The porosity at a given water particle a is calculated by
N
mi
na = ni Wai . (30)
i=1
ρi
By applying Eq. (27) to the continuity equation of water, i.e., Eq. (2), the SPH
form for the continuity equation of water can be obtained as
M M
dρa mb
= mb v ab · ∇a Wab + δf ha cf Ψab · ∇a Wab , (31)
dt ρb
b=1 b=1

where v ab = v a − v b . δf is a constant normally set to 0.1, ha is the smoothing


length of water particle a, cf is the sound speed of water. Since numerical pressure
oscillation occurs in weakly compressible fluid analysis, numerical techniques are
needed. The so-called δ − SPH [Antuono et al. (2010); Marrone et al. (2011)] and
δplus − SPH [Sun et al. (2017)] method could be used to avoid density fluctuation
and thus yield smoother pressure distribution. In this study, the former one was
chosen.Ψab is defined as follows:
xab
Ψab = 2(ρb − ρa ) . (32)
|xab |2
By applying Eq. (28) to Eq. (6), the SPH form of momentum conservation equation
for soil is given as
N

dv i σ i σ j
= mj + 2 − Πij I · ∇i Wij
dt j=1
ρ2i ρj

M M
ma p̃a ma f ia
− (1 − ni ) ∇i Wai + Wai + g i , (33)
ρ ρi
a=1 a
ρ ρi
a=1 a

where σ  i is the effective stress tensor of soil particle, p̃a is the pressure of water
particle. σ  i and p̃a are calculated according to the constitutive relationship of soil
and water, respectively.

1850135-8
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

Similarly, the momentum equation for water can be rewritten in SPH form as
M   M  
dv a p̃a na p̃b nb τ a na τ b nb
=− mb + + Πab ∇ W
a ab + m b + · ∇a Wab
dT ρ2a ρ2b ρ2a ρ2b
b=1 b=1

N N
mi p̃a mi f ia
+ ni ∇a Wai − Wai + g a , (34)
i=1
ρi ρa i=1
ρi ρa

where na is porosity at water particles, calculated using Eq. (30).


In Eqs. (33) and (34), Πij and Πab are the artificial viscosities used for soil
and water, respectively. According to Monaghan [1994], the dissipative term Πij
has been proved to improve the numerical stability and remove the unphysical
penetration between particles. Of the several types of artificial viscosity, the most
widely used one derived by Monaghan is chosen:

 −αΠ cij Φij + βΠ Φ2ij
 , v ij · xij < 0,
Πij = ρij (35)


0, v ij · xij ≥ 0,
where
hij v ij · xij ci + cj
Φij = , cij = , (36)
|xij |2 + 0.01h2ij 2
ρi + ρj hi + hj
ρij = , hij = . (37)
2 2
By replacing i, j with a, b, Πab can be defined analogously. In the above equations,
the two constants αΠ and βΠ are, respectively, set to 0.1 and 1.0 for soil, while
for water they take values of 0.01 and 1.0. αΠ and βΠ should be determined with
care because this approach introduces a shear viscosity and may lead to unphysical
results.
The SPH formulation for viscous drag force f ia is rewritten as
n2a ρ̃f g(v a − v i )
f ia = . (38)
k
The constitutive relationship for soil in SPH form can be obtained and evaluated
at particle i in the following form:
αβ
= σ  i ω̇iβγ + σ  i ω̇iαγ + 2Gėαβ
αγ γβ
σ̇  i i
 
G  αβ
+ K ε̇γγ αβ
i δi − λ̇i 9K sin ψδ αβ
+  si , (39)
J2
in which
   αβ αβ
3αφ K ε̇γγ
i + (G/ J2 )s i ε̇i
λ̇i = . (40)
27αφ K sin ψ + G

1850135-9
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

The strain and spin rate tensors are approximated according to Eqs. (41) and (42),
respectively.
 
N N
1 mj α ∂Wij mj β ∂Wij 
ε̇αβ =  v + v , (41)
i
2 j=1 ρj ji ∂xβi ρ ji ∂xα
j=1 j i

 
N N
1 mj α ∂Wij mj β ∂Wij 
ω̇iαβ = v − v . (42)
2 j=1 ρj ji ∂xβi ρ ji ∂xα
j=1 j i

For water, the pore pressure and deviatoric stress tensor are calculated by Eqs. (12)
and (10), respectively. The strain rate tensor for water is evaluated at particle a by
M M M

m b ∂W ab m b β ∂Wab 2 m b
εαβ
a = vα + v − v ba · ∇a Wab δ αβ . (43)
ρb ba ∂xβa ρb ba ∂xα a 3 ρb
b=1 b=1 b=1

4. Modeling of Interaction Between Structure and Soil/Water


As stated above, simulating interaction between structure and soil/water particles
requires appropriate treatment of boundary deficiency. Hence, the soil/water parti-
cles located near or on the interface could obtain accurate acceleration. The bound-
aries of rigid have been modeled using (a) ghost particles, (b) fluid particles, (c)
normalizing conditions, (d) boundary particle force and (e) particle-to-particle or
particle-to-surface contact based on momentum equations [Monaghan and Kajtar
(2009)]. The pity is that all the above methods are only suit able for handling per-
fectly smooth or nonslip boundaries. Later, a more attractive method that is capable
of dealing with frictional sliding contact was developed by Wang et al. [2013] and
Wang and Chan [2014]. The frictional sliding contact algorithm is actually a penalty
method based on the conservation of momentum, in which the contact force is calcu-
lated according to the amount of partial penetration. Precisely modeling frictional
contact behavior between the structure and soil/water particles is of great impor-
tance in this study. Therefore, the frictional sliding contact algorithm is chosen.
Here, we introduce the final form of the contact force, more details could be found
in [Wang et al. (2013)].
 
2mi
F n = (1 − ς) (d0 + G · n) n, (44)
(∆t)2

 ξ|F n | 
 F τ , if |F  τ | > ξ|F n |,
|F  |
Fτ = τ (45)

 
F τ, otherwise,

2mi
F τ = (∆u − ∆u · n), (46)
(∆t)2
where F n is the normal component of the contact force ς defines the extent of
penetration allowed and was taken as 0.01 ∼ 0.1 throughout this study. The tangent

1850135-10
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

component F τ is calculated through Eq. (45), which is in fact a corrected form.


F  τ is the tangent contact force calculated through the relative velocity when no
sliding is permitted. ξ is the frictional coefficient. G represents the vector from the
coordinate of the particle to its perpendicular foot on the rigid structure. n refers
to the outward normal vector.

5. Numerical Issues and Time Integration


When applied to solid mechanics, SPH may encounter serious tensile instability
problem. It is a particular problem that SPH particles form clumps when material
is stretched. This phenomenon is so severe that in some cases it may lead to com-
pletely wrong results. Swegle et al. [1995] first studied the instability in detail and
related it to the sign of the pressure and the sign of the second derivative of the
kernel function. It has been reported that the tensile instability could be removed by
using hyperbolic kernel function [Yang and Liu (2012); Yang et al. (2014)] and Ten-
sile Instability Control (TIC) method [Sun et al. (2018)]. However, in this study, it
was found that the hyperbolic kernel function couldn’t effectively solve the problem
for water when extremely large soil deformation was encountered. We apply herein
the artificial stress (in the case of fluid, i.e., artificial pressure) method proposed
by Monaghan [2000] and Gray et al. [2001]. In the previous research on simulat-
ing soil–water interaction by SPH, few works about tensile instability treatment
have been mentioned. Bui et al. [2008] demonstrated that artificial stress is only
needed for cohesive soil but not necessary for noncohesive soil. In this study, it
has been found that the fluid phase encounters serious tensile instability when soil
undergoes large deformation. Despite the fact that artificial stress method is not
needed for noncohesive soil, the water particles may clump unphysically, especially
in the case of low hydraulic conductivity. Hence, the artificial pressure method was
implemented for fluid phase in this study. The momentum equation for water, i.e.,
Eq. (34), is replaced by
M  
dv a p̃a na p̃b nb
=− mb + + Πab + f ϑ
(Ra + Rb ) ∇a Wab
dt ρ2a ρ2b ab
b=1

M  
τ a na τ b nb
+ mb 2
+ 2 · ∇a Wab
ρa ρb
b=1

N N
mi p̃a mi f ia
+ ni ∇a Wai − Wai + g a , (47)
i=1
ρi ρa i=1
ρi ρa
where fab is the repulsive force term and specified in terms of the kernel:
Wab
fab = (48)
W (∆d, h)
where ∆d denotes the initial particle spacing. ϑ is usually taken as W (0, h)/
W (∆d, h). For Wendland kernel, n has the value about 3.24 with h equals to 1.2 ∆d.

1850135-11
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

Fig. 1. Leap–Frog time integration scheme.

The factors Ra and Rb are determined in terms of pressure:



 ε|p̃ |
 a , if p̃a < 0,
Ra = ρ2a (49)


0, otherwise,

where ε is a small constant and taken as 0.2 typically. Rb could be obtained by


simply replacing a with b in Eq. (49).
The second-order Leap–Frog (LF) algorithm was chosen to perform time step-
ping [Liu and Liu (2004)]. The LF algorithm has proven to be efficient and requires
low memory. The field variables Θ, i.e., velocity v, density ρ, stress tensor σ αβ and
porosity n, are then integrated at each particle and then the coordinates of particles
are updated. The LF algorithm was performed as follows: Firstly, at the end of the
first time step t0 , field variables Θ leap from t0 to t0 + ∆t/2 whereas coordinate
x leaps from t0 to t0 + ∆t. Secondly, field variables Θ are predicted and leap to
t0 + ∆t aiming at keeping the calculations consistent. Thirdly, the field variables
Θ and coordinate x are updated in the standard LF scheme. A brief schematic
diagram is as follows:
The stability condition of the LF scheme is the Courant–Friedrichs–Levy
condition,

∆t ≤ Ccour min(h/cs , h/cf ), (50)

where Ccour is the Courant coefficient, taken as around 0.3, cf and cs are the sound
speed of fluid and soil, respectively.

6. Numerical Simulations and Analysis


In this section, totally three numerical tests were conducted to check the applica-
bility of the proposed algorithm. The first, two were plain strain analyses and the
other one was simulated under axisymmetric condition. U-tube seepage was mod-
eled aiming at verifying the correctness of the drag force model. The two-dimension
consolidation was designed to test the ability to predict the pore water pressure.
The soil-structure interacting algorithm has been proved to be robust and efficient,
for details see [Wu et al. (2017)]. Finally, the whole continuous penetrating pro-
cess of a spudcan, which involves large soil deformation and soil–water-structure
interaction, was simulated. Since the water is considered as weakly compressible, in

1850135-12
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

the subsequent simulations, the water density is initialized according to the hydro-
static pressure of incompressible fluid. The initial density of the water was inversely
calculated through Eq. (12). For soil, the initial stress state is determined by the
coefficient of lateral earth pressure at rest.

6.1. U-tube seepage modeling


The flow of water in a U-tube from left to right through soil under gravity was
simulated. The aim of the flow simulation was to validate the frictional contact
algorithm and the viscous drag force model. The geometry of this example is shown
in Fig. 2.
The material parameters used and numerical settings are as follows: the ini-
3
tial porosity of soil n = 0.45, particle density of soil ρ̃s = 2,670 kg/m , hydraulic
3
conductivity k = 0.01 m/s, initial intrinsic density of water ρ̃f 0 = 1,000 kg/m ,
sound speed cf = 20 m/s, dynamic viscosity of water µ = 0.001 Pa · s, time step
∆t = 1.0 × 10−5 s. Totally, 2,940 water particles and 400 soil particles are used.
Initial particle spacing ∆d is taken as 0.05 m and the simulation lasts for 200 s. Dif-
ferent from the previous studies, this test uses the frictional contact algorithm to
model the U-tube instead of ghost or virtual particles. The frictional coefficient for
water is taken as 0 to model the free-slip condition. For soil, the frictional coefficient
is assumed to be very large to prevent the soil shifting. It’s worth noting here that
the density diffusive term, i.e., the second term on the right-hand side of Eq. (31),
was not included in this test. Besides, Shepard filtering [Liu and Liu (2004); Yang
and Liu (2013)] method was adopted aiming at improving the accuracy on the free
surface.
As shown in Fig. 3, the flow of water through saturated soil was successfully
simulated. Over time, the water level difference decreased gradually. It could be
seen that at the right-hand side of the soil, the particle spacing is less than the
initial particle spacing. These SPH water particles are initially from the mixture and
the masses of these particles are smaller than those that are not from the mixture.

1.0

Water level
difference Z 1.0
water
3.35
1.0
2.0
soil

4.0

Fig. 2. Geometry of the U-tube seepage model.

1850135-13
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

(a) t = 0 s (b) t = 36 s (c) t = 45 s (d) t = 200 s

Fig. 3. Flow of water through saturated soil in a U-tube at different time.

Fig. 4. Water level difference versus time: Comparison between SPH mixture model and Darcy’s
law.

Smaller particle spacing leads to equivalent density to that of pure water. Figure 4
shows the comparison between the numerical calculated water level difference and
analytic results according to Darcy’s law. Darcy’s law gives the water level difference
δZ at time t as follows:

δZ0
δZ = 2k(t−t0 )
, (51)
e L

where δZ0 is the initial water level difference, t0 is the initial time, L is the length
of soil sample and k is the hydraulic conductivity.
It is shown that there is good agreement between the numerical and analytical
results. Besides, the frictional contact algorithm could effectively prevent soil and
water particles from penetrating the rigid boundary. Since no ghost particles or
boundary particles are needed, the contact algorithm is more efficient.

1850135-14
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

Fig. 5. Geometry of the 2-D consolidation model.

6.2. 2-D consolidation problem


A 2-D consolidation problem was simulated to validate the pressure calculation of
the proposed model. Boer et al. [1993] and Breuer [1999] studied the same problem
by FEM. The parameters used here were taken for comparison. The geometry is
shown in Fig. 5. The soil–water mixture was 20.0 m long and 10.0 m wide, with
15 KPa uniformly distributed load at the top. The left, right and the bottom were
fixed and undrained, whereas the top was drained. The soil parameters are: λs =
3
5,583 KPa, µs = 8,375 KPa, ρ̃s = 2,000 kg/m , n = 0.33, υ = 0.2, k = 0.01 m/s.
3
The water parameters are: ρ̃f = 1,000 kg/m , cf = 97.5 m/s. Totally, 800 particles
were used with initial resolution ∆d = 0.5 m. The left, right and bottom were all
modeled as nonslip using ghost particles. The initial numerical model is shown in
Fig. 6. Since the mixture theory was used here, the loading ratio of the soil and
water was indeterminate. A novel method was developed here to simulate the stress
boundary condition at the top.

Fig. 6. Initial SPH model of 2-D consolidation: Boundary of saturated soil modeled by ghost
particles; stress boundary modeled by contact algorithm.

1850135-15
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

The rigid boundary at the top was assigned velocity through


αcs Acs ∆t
n
vR =   (σg − σm ), (52)
Ns Nf
2 i=1 mi + a=1 ma

where Acs is the contact area. Ns and Nf are the number of soil and water particles
contacting with the rigid boundary. mi and ma are the mass of soil and water
particles, respectively. σg is the target value of contact stress, that is 15 KPa for
this test. αcs is a scaling factor used to weaken the oscillation, taken as 0.01, σm is
the measured contact stress in present time step, given by
Ns Nf
i=1 Fi + a=1 Fa
σm = , (53)
Acs
where Fi and Fa are calculated through Eq. (44).
The evaluation of the excess pore water pressure at different intervals of time
is shown in Fig. 7. FEM results (left: in KPa) by Breuer [1999] are included for

(a) t = 0.5 s

(b) t = 2 s

Fig. 7. Comparison of excess pore water pressure between FEM (left: in KPa) and SPH (right: in
Pa) at different intervals of time.

1850135-16
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

(c) t = 8 s

(d) t = 10 s

Fig. 7. (Continued.)

comparison. It is shown that the SPH results are consistent with the FEM. In the
beginning (t = 0.01 s), the deformation of soil lagged behind the water, the whole
external load was mainly carried by the water. Hence, the excess water pressure
increased to around 14,000 Pa quickly. With the passage of time, the soil skeleton
carried more external load and the excess pore pressure decreased gradually. After
10 s, the excess pore pressure was about zero. It was proved that the proposed
SPH mixture model could evaluate the excess pore water pressure satisfactorily.
Moreover, the servo control method modeling the stress boundary was efficient.

6.3. Spudcan penetration simulation


Spudcan is an inverted conical structure used to support mobile jack-up rigs in
offshore drilling. The installation of the spudcan usually involves extremely large soil
deformation and complex soil–water-structure interaction [Hossain and Randolph
(2010)]. Few numerical research works have been reported on the whole continuous
penetration process, meanwhile, considering the effect of pore water.

1850135-17
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

In this section, the continuous penetration process of the spudcan was simulated
using SPH based on the proposed soil–water-structure interaction model. Two sim-
ulation cases, i.e., sand and clay, were conducted under axisymmetric conditions. It
should be noted here that the governing equations should be modified for this test to
fit the cylindrical coordinates. Besides, the symmetry axis was treated by combining
the ghost particle method and the boundary contact algorithm in Sec. 4. Additional
numerical techniques were also needed to prevent the singularity from happening.
For more details about axisymmetric SPH algorithm, refer to Wang et al. [2016b].
The schematic diagram of the numerical model is shown in Fig. 8. Both soil samples
were immersed in water with free surface. The bottom and right boundaries were all
fixed and undrained. The left boundary, i.e., the axis of symmetry, was free-slip. All
boundaries were modeled using ghost particles except for the spudcan. The spudcan
was rigid and moved downward at a given speed leading to a quasi-static state. The
interaction between the spudcan and the mixture was simulated using the frictional
sliding contact algorithm. Drucker–Prager model with nonassociated flow rule was
adopted for both sand and clay. It should be noted here the objective of this paper
was to propose a computational approach. Therefore, the relatively simple consti-
tutive model was chosen. The material properties, the calculation parameters and
the geometry are shown in Table 1.

Fig. 8. Schematic diagram of the spudcan penetration model.

Table 1. Simulation cases: soil, water parameters and geometry.


Case 1 Sand (medium dense) ρ̃s = 2,650 kg/m3 , ID = 0.5, Es = 30 MPa, n = 0.44,
k = 1.0 × 10−5 m/s, υ = 0.3, c = 0 kPa, φ = 31.5◦ , ψ = 4.5◦
Water ρ̃f 0 = 1,000 kg/m3 , cf = 97.5 m/s, µ = 0.001 Pa · s
Geometry Hw = 3 m, Hsw = 7 m, Wsw = 12 m, Dsp = 6 m,
Lb = 0.54 m, Lm = 0.15 m, Lt = 0.858 m
Case 2 Kaolin clay ρ̃s = 2,600 kg/m3 , Es = 6.5 MPa, c = 1 kPa, kE = 520 kPa/m,
k = 2.0 × 10−7 m/s, n = 0.63, υ = 0.33, φ = 23◦ , ψ = 0◦
Water ρ̃f 0 = 1,000 kg/m3 , cf = 142.0 m/s, µ = 0.001 Pa · s
Geometry Hw = 6.25 m, Hsw = 15.0 m, Wsw = 25.0 m,
Dsp = 12.5 m, Lb = 2.0 m, Lm = 1.0 m, Lt = 1.5 m

1850135-18
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

Fig. 9. Penetration process of the spudcan into sand (medium dense).

Figure 9 shows the continuous penetration process of the spudcan into medium
dense sand. The spudcan was initially underwater while above the soil surface. As
the spudcan penetrated, a part of the soil beneath the bottom was forced downwards
while a small amount of soil at the edge flows upward and forms the heave. The
initially superimposed soil and water particles moved separately. As the penetration
continues, the cavity forms above the spudcan and remains open. The back-flow of
soil doesn’t occur. It was because the maximum penetration depth was only 4 m.
The water level kept unchanged. It could be seen that by using SPH the free surface
of both water and soil could be conveniently tracked without any other techniques.
The penetration resistance during installation is shown in Fig. 10. The CEL pre-
dicted results by Qiu and Grabe [2012] is also included for comparison. The SPH
results show good agreement with CEL results before the normalized penetration
depth reaching 0.3. The peak value could be observed at a normalized depth of
0.3. The maximum penetration resistance was around 550 kPa. SPH results with
or without considering the soil porosity showed no significant differences. Further

Fig. 10. Comparison of normalized penetration depth versus penetration resistance predicted by
CEL and SPH (with and without considering of porosity).

1850135-19
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

penetration leads to a reduction of penetration resistance predicted by CEL. The


reason is that the CEL results are based on penetrating into two-layer soil stratum,
where there is soft clay underneath the sand. However, single sand layer was sim-
ulated here for simplicity. This does not detract from the objective of this study,
which is to propose a computational approach. As a result, only the initial pene-
tration stage, i.e., at a normalized depth of around 0.35, is needed to validate the
proposed method. It was worth noting that as the penetration continued the results
predicted by the proposed method are 5% lower than without considering the soil
porosity.
As mentioned above, to date, there are few numerical studies on the excess
pore pressure around a penetrating spudcan. In fact, when dealing with problems
involving failure and large deformation, to compute excess pore pressure using FEM
is difficult and rarely reported. Yi et al. [2012] conducted the undrained effective
stress analysis by separately defining the soil skeleton’s effective stress-strain matrix
and bulk modulus of water and superimposing the two components in the user-
defined material based on ABAQUS/Explicit. This only numerical study on the
generation and dissipation of excess pore pressure around the spudcan by Yi et al.

(a) d/D = 0.4

(b) d/D = 0.8

Fig. 11. Comparison of excess pore water pressure at normalized depths of 0.4 and 0.8 between
effective-stress analysis (left: in KPa) and coupled soil–water interaction analysis (right: in Pa).

1850135-20
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

[2012] is an approximate treatment and does not take the soil–water interaction
into account. Figures 11(a) and 11(b) show excess pore water generated beneath
the spudcan at the normalized depth of 0.4 and 0.8 calculated by the proposed
method (on the right). Results by Yi et al. [2012] are also shown (on the left) for
comparison. At shallow penetration depth, i.e., d/D = 0.4, the maximum excess
pore pressure was around 60 kPa directly below the spudcan. As the normalized
penetration depth reached 0.8, the maximum excess pore pressure increased to
130 KPa. Comparison of the left and right of Figs. 11(a) and 11(b) shows that
the excess pore pressure computed by SPH tends to decrease more rapidly with
the radial distance. As shown in Fig. 11(b), the excess pore pressure bulb extends
laterally outwards to about 12.0 m. In contrast, the left reaches around 17.0 m. In
the vicinity of the edge of the spudcan, the excess pore pressure appears to be
irregular. This may be attributed to the fact that the soil near the edge undergoes
extremely large shear deformation whereas there are not enough SPH particles.
In the future studies, in order to improve the accuracy while avoiding too much
calculation, adaptive particle refinement techniques [Wu et al. (2017); Sun et al.
(2018)] will be employed in the areas undergoing large deformation.

7. Conclusions
SPH has been extensively utilized for its mesh-free nature. Large soil deformation
and soil–water-structure interaction are commonly seen in geotechnical, coastal and
ocean engineering, which are difficult to simulate using mesh-based methods.
A fully coupled soil–water-structure interaction algorithm has been proposed in
the framework of SPH. In this algorithm, soil–water interaction was depicted by two-
phase mixture theory. Each phase of the mixture occupies part of the macroscopic
mixture and satisfies its own conservation equations of mass and momentum. The
interaction force between two phases consists of pore pressure and viscous drag
force, the latter of which was governed by Darcy’s law. The constitutive behavior
of soil was described using Drucker–Prager model with nonassociated plastic flow
rule and the water was considered as Newtonian fluid. Each phase of the mixture
was represented by different types of SPH particles and moved according to its
own governing equations. The structure was assumed to be rigid. The interaction
between each phase of the mixture and the structure was treated separately based
on frictional sliding contact algorithm.
Firstly, the viscous drag model was validated by the simulation of flow of water
in a U-tube through soil under gravity. The variation of water level difference was
compared to theoretical results and good agreement was obtained. Secondly, a two-
dimensional consolidation problem was simulated to validate the calculation of the
excess pore water pressure. A novel method was proposed here to simulate the
boundary condition of constant stress. It was shown that by using the mixture model
the generation and dissipation of the excess pore water pressure could be efficiently
captured. Thirdly, the penetration process of the spudcan, which involves large soil

1850135-21
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

deformation and complex interaction between the mixture and the spudcan, was
simulated under axisymmetric conditions. Both penetration resistance and excess
pore water pressure were investigated and compared to the results calculated using
LDFE method.
The above numerical simulations showed that with the proposed soil–water-
structure interacting model, (1) the large deformation of soil could be conveniently
dealt with; (2) the excess pore water pressure could be efficiently obtained; (3) the
effect of soil porosity could be taken into account; (4) the frictional resistance acting
on the structure could be satisfactorily predicted.

Acknowledgments
This research is supported by the National Natural Science Foundation of China
(Grant Nos. 41727802, 51678360, and 51779084). Also, the authors want to thank
the corrections and constructive suggestions made by the referee, which improved
the presentation of this paper.

References
Antuono, M. et al. [2010] “Free-surface flows solved by means of SPH schemes with numer-
ical diffusive terms,” Comput. Phys. Commun. 181, 532–549.
Bagnold, R. A. [1954] “Experiments on a gravity-free dispersion of large solid spheres in
a Newtonian fluid under shear,” Proc. Roy. Soc. Lond. Ser. A Math. Phys. Sci. 225,
49–63.
Biot, M. A. [1941] ‘General theory of three-dimensional consolidation,” J. Appl. Phys. 12,
155–164.
Boer, R. D. et al. [1993] “One-dimensional transient wave propagation in fluid-saturated
incompressible porous media,” Arch. Appl. Mech. 63, 59–72.
Breuer, S. [1999] “Quasi-static and dynamic behavior of saturated porous media with
incompressible constituents,” Transp. Porous. Media. 34, 285–303.
Bui, H. H. et al. [2007] “Numerical simulation of soil–water interaction using smoothed
particle hydrodynamics (SPH) method,” J. Terramech. 44, 339–346.
Bui, H. H. et al. [2008] Lagrangian mesh-free particle method (SPH) for large deformation
and post-failure of geomaterial using elastic-plastic soil constitutive model, Int. J.
Numer. Anal. Methods. Geomech. 32, 1537–1570.
Bui, H. H. and Fukagawa, R. [2013] “An improved SPH method for saturated soils and its
application to investigate the mechanisms of embankment failure: Case of hydrostatic
pore-water pressure,” Int. J. Numer. Anal. Methods. Geomech. 37, 31–50.
Coussy, O. et al. [1998] “From mixture theory to Biot’s approach for porous media,” Int.
J. Solids Struct. 35, 4619–4635.
Drew, D. A. [1983] “Mathematical modeling of two-phase flow,” Annu. Rev. Fluid. Mech.
15, 261–291.
Gingold, R. A. and Monaghan, J. J. [1977] “Smoothed particle hydrodynamics: Theory
and application to non-spherical stars,” Mon. Not. Roy. Aston. Soc. 181, 375–389.
Gray, J. P. et al. [2001] “SPH elastic dynamics,” Comput, Meth. Appl. Mech. Eng. 190,
6641–6662.
Hossain, M. S. and Randolph, M. F. [2010] “Deep-penetrating spudcan foundations on
layered clays: Numerical analysis,” Geotechnique. 60, 171–184.

1850135-22
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

Soil–Water-Structure Interaction Algorithm in Smoothed Particle Hydrodynamics

Huang, Y. et al. [2013] “Numerical simulation of flow processes in liquefied soils using
a soil–water-coupled smoothed particle hydrodynamics method,” Nat. Hazards. 69,
809–827.
Iverson, R. M. and Denlinger, R. P. [2001] “Flow of variably fluidized granular masses
across three-dimensional terrain 1. Coulomb mixture theory,” J. Geophys. Res. 106,
537–552.
Liu, G. R. and Liu, M. B. [2004] Smoothed Particle Hydrodynamics: A Meshfree Particle
Method (World Scientific: Singapore).
Liu, M. B. et al. [2008] “An overview on smoothed particle hydrodynamics,” Int. J. Com-
put. Methods 5, 135–188.
Liu, M. B. and Li, S. M. [2016] “On the modeling of viscous incompressible flows with
smoothed particle hydrodynamics,” J. Hydrodyn. 28, 731–745.
Liu, M. B. and Liu, G. R. [2010] “Smoothed Particle Hydrodynamics (SPH): An overview
and recent developments,” Arch. Comput. Method Eng. 17, 25–76.
Marrone, S. et al. [2011] “δ-SPH model for simulating violent impact flows,” Comput.
Meth. Appl. Mech. Eng. 200, 1526–1542.
Monaghan, J. J. [1994] “Simulating Free Surface Flows with SPH,” J. Comput. Phys. 110,
399–406.
Monaghan, J. J. [2000] “SPH without a tensile instability,” J. Comput. Phys. 159, 290–
311.
Monaghan, J. J. and Kajtar, J. B. [2009] “SPH particle boundary forces for arbitrary
boundaries,” Comput. Phys. Commun. 180, 1811–1820.
Qiu, G. and Grabe, J. [2012] “Numerical investigation of bearing capacity due to spudcan
penetration in sand overlying clay,” Can. Geotech. J. 49, 1393–1407.
Sun, P. N. et al. [2017] “δ plus-SPH model: Simple procedures for a further improvement
of the SPH scheme,” Comput. Meth. Appl. Mech. Eng. 315, 25–49.
Sun, P. N. et al. [2018] “Multi-resolution SPH with tensile instability control: Towards
high Reynolds number flows,” Comput. Phys. Commun. 224, 63–80.
Swegle, J. W. et al. [1995] “Smoothed Particle Hydrodynamics Stability Analysis,” J.
Comput. Phys. 116, 123–134.
Takeda, H. et al. [1994] “Numerical simulation of viscous flow by smoothed particle hydro-
dynamics,” Prog. Theor. Phys. 92, 939–960.
Terzaghi, K. [1943] Theoretical Soil Mechanics (Wiley, New York).
Wang, C. et al. [2016a] “Smoothed Particle Hydrodynamics Simulation of Water-Soil Mix-
ture Flows,” J. Hydraul. Eng.-ASCE. 142, 04016032.
Wang, J. et al. [2016b]. “Stable axisymmetric SPH formulation with no axis singularity,”
Int. J. Anal. Methods Geomech. 40, 987–1006.
Wang, J. et al. [2013] “Simulating frictional contact in smoothed particle hydrodynamics,”
Sci. China-Technol. Sci. 56, 1779–1789.
Wang, J. and Chan, D. [2014] “Frictional contact algorithms in SPH for the simulation of
soil–structure interaction,” Int. J. Anal. Methods Geomech. 38, 747–770.
Wang, Y. and Hutter, K. [1999] “A constitutive model of multiphase mixtures and its
application in shearing flows of saturated solid-fluid mixtures,” Granul. Matter. 1,
163–181.
Wu, H. et al. [2017] “Asymmetric adaptive particle refinement in SPH and its application
in soil cutting problems,” Int. J. Comput. Methods. 15, 1850052.
Yang, X. F. and Liu, M. B. [2012] “Improvement on stress instability in smoothed particle
hydrodynamics,” Acta Phys Sin-ch Ed. 61, 224701.
Yang, X. F. and Liu, M. B. [2013] “Numerical modeling of oil spill containment by boom
using SPH,” Sci. China Phys. Mech. 56, 315–321.

1850135-23
2nd Reading
July 19, 2018 15:24 WSPC/0219-8762 196-IJCM 1850135

H. Wu et al.

Yang, X. F. et al. [2014] “Smoothed particle hydrodynamics modeling of viscous liquid


drop without tensile instability,” Comput. Fluids. 92, 199–208.
Yi, J. T. et al. [2012] “Eulerian finite element analysis of excess pore pressure generated
by spudcan installation into soft clay,’ Comput. Geotech. 42, 157–170.
Zienkiewicz, O. C. and Shiomi, T. [1984] “Dynamic behavior of saturated porous media:
The generalized Biot formulation and its numerical solution,” Int. J. Numer. Anal.
Methods. Geomech. 8, 71–96.

1850135-24

View publication stats

You might also like