You are on page 1of 13

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

net/publication/318275177

Bearing capacity bearing capacity of shallow foundation by smoothed particle


hydrodynamics (SPH) analysis

Conference Paper · April 2011

CITATIONS READS

6 1,877

2 authors:

Ha H. Bui Dat Vu Khoa Huynh


Monash University (Australia) Norwegian Geotechnical Institute
142 PUBLICATIONS   3,565 CITATIONS    29 PUBLICATIONS   485 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Management of coal bursts and pillar burst in deep coal mines View project

Continuum-discrete modelling of damage and fracture in engineering materials View project

All content following this page was uploaded by Ha H. Bui on 07 July 2017.

The user has requested enhancement of the downloaded file.


Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

BEARING CAPACITY OF SHALLOW FOUNDATION BY SMOOTHED


PARTICLE HYDRODYNAMICS (SPH) ANALYSIS

H.H. Bui
Department of Civil Engineering, Ritsumeikan University, Japan
H.D.V. Khoa
Norwegian Geotechnical Institute (NGI), Norway

ABSTRACT: The paper illustrates the potential of Smoothed Particle Hydrodynamics


(SPH) computing technique for accurate and efficient analyzing of bearing capacity and
failure mechanism of shallow foundation. SPH is a meshfree particle method based on
Lagrangian formulation. Since the SPH method does not possess any mesh-related
difficulties, it is more suitable to deal with large geometric changes of the domain of interest
than the traditional mesh-based discretisation techniques such as finite elements, finite
differences or finite volumes. Although the advantages of using SPH in the field of fluid and
solid mechanics have been well described by many authors, its application in geotechnical
engineering is still very limited. As an application of the SPH method to geotechnical
problems, deep penetration of a shallow foundation into soil is simulated under plane strain
conditions. The soil behavior is modelled by the non-associated Drucker-Prager model
implemented into the SPH code. The computed results in terms of bearing capacity and
failure mechanism are compared with the PLAXIS finite element results and the Terzaghi
limit equilibrium method. The good agreement in the results emphasizes the potential
application of the SPH method to a wide range of geotechnical problems.

1 INTRODUCTION
Smooth Particle Hydrodynamics (SPH), a meshfree particle method based on Lagrangian
formulation, was introduced in 1977 (Lucy, 1977; Gingold and Monaghan, 1977). It has been
successfully applied to astrophysical problems and later widely extended to solve the problem
of continuum solid mechanics and fluid mechanics. The SPH method discretizes the
continuum body only with a set of nodal points (particles), and so it does not possess any
mesh-related difficulties. In this method, the partial differential equations for the continuum
are converted into equations of motion of these particles and then solved by updated
Lagrangian numerical scheme. Compared with other traditional mesh-based discretisation
techniques such as finite element, finite differences or finite volumes, SPH has several
advantages as follows:
- it can handle large deformation and post-failure very well thanks to its Lagrangian
particle-based format and adaptive nature;
- complex free surfaces are modeled naturally without any special treatments;
- complex geometries can be handled without any difficulties and it is relatively easy to
incorporate complicated physics.
Recently, Bui and his co-workers have been intensively focused on the development an
SPH model for modelling large deformation problems and failure flows in geomaterials (Bui

457
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

et al., 2007; 2008; 2010). Many important geotechnical problems including granular flows,
bearing capacity, slope stability, soil-water coupling, soil-structure interaction, seepage
flows…etc, have been successfully analyzed using the SPH method. As an example of the
application of the SPH method to geotechnical problems, this paper discusses the bearing
capacity and failure mechanism analysis of deep penetration of a shallow foundation into soil.
The elasto-plastic soil behavior is modelled by the Drucker-Prager model implemented into
the SPH code. The computed results in terms of bearing capacity and failure mode are
compared with the PLAXIS finite element (FE) solutions to validate and highlight the SPH
based approach.
The rest of the paper is organized as follows. Section 2 briefly describes the smoothed
particle hydrodynamics method. In Section 3, the SPH discretisation of motion equation is
recalled and the technique used for improving the numerical stability is also discussed here.
Section 4 focuses on description of the elastic-plastic Drucker-Prager constitutive model. The
equations used for matching the Drucker-Prager model and the Mohr-Coulomb model are
also presented in this section. Section 5 assesses the ability of numerical analysis using the
SPH method to predict the bearing capacity and failure mechanism of a shallow foundation.

2 SMOOTHED PARTICLE HYDRODYNAMICS


In SPH, approximations for quantities of a continuum field such as density, velocity,
pressure, etc., are approximated using the following integral function,

A(r ) = ∫ A(r')δ ( r − r ' ) dr ' (1)

where A is any variables defined on the spatial coordinate r, δ refers to the Direct delta
function. This integral is then approximated by replacing the delta function with a smoothing
kernel W with characteristic h, such that

lim W (r − r ', h) = δ (r − r ') (2)


h→0

giving
A(r ) = ∫ A(r ')W ( r − r ' , h) dr '+ O ( h 2 ) (3)

The kernel function is normalized according to,

Ω
∫ W (r − r ', h) dr ' = 1 (4)

The choice of kernel function directly affects the accuracy, efficiency and the stability of
numerical algorithm. A number of kernel functions have been proposed in the SPH literature
so far, we apply herein the most commonly used kernel function, namely cubic-spline
function proposed by Monaghan & Lattanzio (1985), which has the following form,

⎧ 1 − 32 q 2 + 34 q 3 0 ≤ q <1

W (q ) = α d × ⎨ 14 (2 − q )3 1≤ q < 2 (5)
⎪0 q≥2

where αd is the normalization factor which is 10/7πh2 in two-dimensional problems and q is


the normalized distance q=|r|/h.

458
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

Finally, the integral (3) is discretised onto a finite set of interpolation points (the particles)
by replacing the integral with a summation and the mass element ρV with a particle mass m,
i.e.

A(r ')
A(r ) = ∫ W ( r − r ' , h) ρ (r ')dr '+ O(h 2 )
ρ (r ')
N
(6)
Ab
≈ ∑ mb W ( r − rb , h)
b =1 ρb

where subscript b refers to the quantity evaluated at the position of particle b. This
“summation approximation” is the basis of all SPH formalisms.
The SPH approximation for the gradient terms may be calculated by taking analytical
derivative of Eq. (6), giving:

∂ A(r ')
∇A(r ) = ∫
∂r ρ (r ')
W ( r -r ' , h) ρ (r ')dr '+ O(h 2 )
N
(7)
Ab
≈ ∑ mb ∇ aWab
b =1 ρb

where we have assumed that the gradient is evaluated at another particle a (r = ra) and the
remaining terms are defined by,

rab ∂Wab
Wab ≡ W ( ra − rb , h) and ∇ aWab ≡ (8)
rab ∂ra

However, this form of gradient is not guaranteed to vanish when A(r) is not constant. To
ensure that it does, the gradient can be written as,
N
( Ab − Aa )
∇Aa ≈ ∑ mb ∇ aWab (9)
b =1 ρb

which is an estimate of

∇A(r ) = ∇A − A(∇1) (10)

Alternatively, the following forms of gradient approximation which are the most
commonly used to discrete the momentum equation can be written as (Monaghan, 1992;
Bonet et al., 1999),

N
⎛A A ⎞
∇Aa ≈ ρ a ∑ mb ⎜ a2 + b2 ⎟∇ aWab (11)
b =1 ⎝ ρ a ρb ⎠
N
( A + Aa )
∇Aa ≈ ρ a ∑ mb b ∇ aWab (12)
b =1 ρ a ρb

These two last gradient approximations conserve linear and angular momentum
(Monaghan, 1992). Further details of SPH integration scheme as well as other issues of SPH
can be found in Liu & Liu (2003).

459
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

3 SPH DISCETIZATION OF MOTION EQUATION


The motion of a continuum can be described through the following equation,

ρ uα = ∇ β σ αβ + ρ g α (13)

where u is the displacement; α and β denote Cartesian components x, y, z with the Einstein
convention applied to repeated indices; ρ is the density; σ is the total stress tensor, taken
negative for compression; and g is the acceleration due to gravity. For a soil, the total stress is
normally composed of the effective stress tensor (σ′ ) and the pore-water pressure (pw),

σ αβ = σ ′αβ + pwδ αβ (14)

In the current study, the pore-water pressure is not considered, thus the displacement u of
soil particles relates to the effective stress in the following way,

ρ uα = ∇ β σ ′αβ + ρ g α (15)

Using Eq. (11), the partial differential form of Eq. (15) can be approximated in the SPH
formulation in the following ways,

N
⎛ σ ′αβ σ ′αβ αβ ⎞ β
uaα = ∑ mb ⎜ a 2 + b 2 + Cab α
⎟ ∇ a Wab + g a (16)
b =1 ⎝ ρa ρb ⎠

where a indicates the particle under consideration; ρa and ρb are the densities of particle a and
b respectively; N is the number of “neighbouring particles”, i.e. those in the support domain
αβ
of particle a; mb is the mass of particle b; Cab is a stabilization term employed to remove the
stress fluctuation and tensile instability (Bui et al., 2008). The stabilization term in Eq. (16)
consists of two components: artificial viscosity and artificial stress, which could be computed
similarly to Bui et al. (2008) except that the sound speed for the artificial viscosity term is
calculated herein by,

ca = Ea / ρ a (17)

where E is Young’s modulus of the soil. Eq. (16) can be integrated using the standard
Leapfrog algorithm if the effective stress tensor is known. Thus, it is necessary to derive a
constitutive relation for the effective stress tensor that is applicable in the SPH framework.

4 SOIL CONSTITUTIVE MODEL


A constitutive model describes the relation between stress and strain of a given material.
Following the classical theory of plasticity, the total strain-rate tensor of an elasto-plastic
material ε is decomposed into two parts: an elastic strain rate tensor ε e and a plastic strain
rate tensor ε p :

εαβ = εeαβ + εαβ


p (18)

460
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

The elastic strain rate tensor εeαβ is given by a generalized Hooke’s law, i.e.,

s′αβ 1 − 2υ m αβ
εe
αβ
= + σ ′ δ (19)
2G E

where s ′αβ is the deviatoric effective shear stress tensor; G is the shear modulus; and σ ′ m is
the mean effective stress.
The plastic strain rate tensor ε αβ
p is calculated by the plastic flow rule,

 ∂g p
εαβ
p =λ (20)
∂σ ′αβ

where λ is the rate of change of plastic multiplier, and gp is the plastic potential function. In
the current study, the Drucker-Prager model with associated and non-associated plastic flow
rules is applied, under the assumptions that the yield surface is fixed in stress space, and
plastic deformation occurs only if the stress state reaches the yield surface. Accordingly,
plastic deformation will occur only if the following yield criterion is satisfied,

f ( I1 , J 2 ) = αϕ I1 + J 2 − kc = 0 (21)

where I1 and J2 are the first and second invariants of the stress tensor; and αφ and kc are
Drucker-Prager constants that are calculated from the Coulomb material constants c
(cohesion) and φ (internal friction). In plane strain, the Drucker-Prager constants are
computed by,

tan ϕ 3c
αϕ = and kc = (22)
9 + 12 tan 2 ϕ 9 + 12 tan 2 ϕ

The associate plastic flow rule takes the same form of the yield function as the plastic
potential function, while the non-associated plastic flow rule specifies the plastic potential
function by,

g p = αψ I1 + J 2 − constant (23)

where αψ is a dilatancy factor which can be related to the dilatancy angle ψ in a fashion
similar to that between αφ and friction angle φ. Substituting Eq. (23) into Eq. (20), and then
Eqs. (19-20) into Eq. (18), and additionally adopting the Jaumman stress rate for large
deformation treatment, the stress-strain relation for the current soil model at particle i
becomes (Bui et al., 2008),

dσ i′αβ
= σ i′αγ ω iβγ + σ i′γβ ω iαγ + 2Gi eiαβ + Kiεiγγ δ iαβ − λi ⎡⎣3Kiαψ i δ αβ + (G / J 2 )i si′αβ ⎤⎦ (24)
dt

461
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

where eiαβ = εiαβ − 13 εiγγ δ αβ is the deviatoric strain rate tensor; λi is the rate of change of
plastic multiplier, which in SPH is specified by,

 3αϕ K εiγγ + (G / J 2 )i si′αβ εiαβ


λi = (25)
9αϕ K iαψ i + Gi

and εiαβ , ω iαβ are the strain rate and spin rate tensors defined by

1 ⎛ ∂uα ∂u β ⎞ 1 ⎡ N m j α ∂Wij N m j β ∂Wij ⎤


εiαβ = ⎜ ⎢∑ + ∑ (u j − uiβ ) α ⎥
α
β
+ α ⎟
= (u − 
u ) (26)
∂x ⎠i 2 ⎢⎣ j =1 ρ j β
j =1 ρ j
j i
2 ⎝ ∂x ∂xi ∂xi ⎥⎦
1 ⎛ ∂uα ∂u β ⎞ 1 ⎡ N m j ∂Wij N m j ∂Wij ⎤
ω iαβ = ⎜ β − α ⎟ = ⎢ ∑ (uαj − uiα ) β − ∑ (u βj − uiβ ) α ⎥ (27)
2 ⎝ ∂x ∂x ⎠i 2 ⎢⎣ j =1 ρ j ∂xi j =1 ρ j ∂xi ⎥⎦

The above soil constitutive model requires six soil parameters, which are cohesion
coefficient (c), friction angle (φ), dilatancy angle (ψ), Young’s modulus (E), Poisson’s ratio
(υ), and soil density (ρ). In the current paper, Eq. (25) is explicitly integrated using the
standard leapfrog algorithm. A numerical error correction method proposed by Chen &
Mizuno (1990) is applied to adjust stresses which exceed the yield strength of material.

5 BEARING CAPACITY ANALYSIS OF A SHALLOW FOOTING


This section discusses the potential of the SPH computing technique for accurate and
efficient analyzing of bearing capacity and failure mechanism of deep penetration of a
shallow foundation into soil. The material behavior is modelled by the elastic-plastic
Drucker-Prager constitutive model described in Section 4. This material model was
implemented into both the SPH code and the commercial finite element (FE) code PLAXIS
(as a user-defined model). The SPH results in terms of bearing capacity and failure mode are
compared with those computed by the FE method for two cases: the associated and non-
associated flow rules. The ultimate bearing capacities predicted by the SPH and FE methods
will be then compared with the well-known conventional bearing capacity solution propose
by Terzaghi (1943).
5.1 Conventional design of bearing capacity using limit equilibrium method
The bearing capacity of shallow foundations is generally calculated using the Terzaghi
method (Terzaghi, 1943), which is based on an approximation solution using superposition
assumption to combine the effect of cohesion, overburden pressure and soil weight. For a
strip footing subjected to vertical and central loading as shown in Fig. 1, the ultimate bearing
capacity qu can be expressed by Terzaghi’s equation as

1
qu = c ⋅ N c + q ⋅ N q + γ ⋅ B ⋅ Nγ (28)
2

where c is the cohesion intercept of soil, q the overburden pressure at foundation depth, γ the
unit weight of soil, B the width of the strip footing, and Nc, Nq and Nc are non-dimensional
bearing capacity factors that are functions of soil friction angle φ.

462
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

Fig. 1 also illustrates the general failure mechanism in soil underlying a shallow strip
footing postulated by Terzaghi based on Prandlt’s theory. There are three distinct failure
zones of soil under the footing: a weight triangular zone (active zone), two Prandtl’s radial
shear zones and two Rankine passive zones.

Applied load

Overburden pressure

Passive 
Radial 
zone
shear 
zone Lowest 
shear 
Soil Wedge zone (active zone) surface

Fig. 1. General failure mechanism in soil at ultimate load for a rigid strip footing (Terzaghi, 1943).

5.2 Numerical analysis of bearing capacity using the SPH and FE methods
The bearing capacity analysis of a shallow layer of clay has been performed using a plane
strain analytical model shown in Fig. 2. Its configuration is similar to those studied by
Zienkiewicz et al. (1975), Chen & Mizuno (1990), except for the width of the model which is
extended to 10 m from the center line. This is to allow the failure to fully develop. The
foundation was assumed to be rigid and in perfect contact with the soil. This means that
relative displacement between the foundation and soil is not permitted. The two vertical sides
of the model have been restrained in the horizontal direction, while the base of the soil layer
was not allowed to move in either the vertical or horizontal direction.

CL

Applied footing velocity (or displacement)

1.57 m

Modelled 3.66 m
half‐symmetry

10.00 m

Fig. 2. Idealization of the plane strain strip foundation problem.

The soil weight effect was neglected and so initial stress state is set at the origin of the
stress space. The Mohr-Coulomb constitutive model with Young’s modulus E = 207 MPa,
Poisson’s ratio ν = 0.3, cohesion c = 69 kPa and internal friction angle φ = 20o was used by
Zienkiewicz et al. (1975) in order to simulate the behaviour of the soil. For plane strain
conditions, the Mohr-Coulomb parameters (φ, c) can be exactly converted to Drucker-Prager
parameters (αφ, kc) using Eq. (22) presented in Section 4.

463
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

• Description of the SPH model


For the SPH model, a total of 9900 SPH particles were arranged in a rectangular lattice to
generate the modelled half-symmetry with an initial smoothing length of 0.0732 m. These
particles have the same material properties as the soil model used in the FEM simulation.
Boundary particles with no-slip boundary condition were used to model the rigid, rough base
and strip footing. On the other hand, the ghost particles that satisfy the free-slip boundary
condition were assigned to the vertical boundary. A speed of sound of 270 m/s determined
based on the Young’s modulus and the density of the soil was used for calculating the
artificial viscosity (see Section 3). This artificial viscosity was then incorporated in the
momentum equation to improve the numerical stability as well as to damp out unwanted
oscillations. The gravitational force was set to zero to model the weightless soil. Even so, a
reference density of 26.5 kN/m3 still need to be assigned to soil particles in order to remove
the numerical divergence from the SPH momentum equation. The foundation load was
applied using a constant downward velocity of 5 cm/s, which is assumed to be small enough
to ensure the accurate solutions for the problem described in this paper.
• Description of the FE model
For the PLAXIS FE-model, 15-nodes triangular elements were used for the soil while the
strip footing was simply modelled by the rigid beam element. The FE model composes of
1314 elements which correspond to 15768 material points (i.e. Gauss points). The foundation
load was simulated by applying increments of vertical displacement to the rigid beam.
5.3 Comparison of the bearing capacity results
The bearing capacity responses of the rigid strip footing calculated by using the SPH method
were compared with those calculated using the PLAXIS FE method and the Terzaghi limit
equilibrium solution.
Fig. 3 plots the load-displacement responses the strip footing for both the associated and
non-associated flow rules (zero dilatancy angle). It can be seen from the figure that the
bearing capacities obtained with the SPH model in this paper (named current SPH model) are
significantly small than those obtained with the same SPH simulation presented in the paper
of Bui et al. (2008) (named SPH model 2008). This can be explained by three main effects:
- Effect of the SPH discretisation: the number of particles in the current model is larger
than in the model 2008 (9900 SPH particles compared to 7371 SPH particle). This
finer model can minimize the overshoot of the calculated bearing capacity caused by
the “coarse” SPH discretisation used in the SPH model 2008.
- Effect of boundary: the load-displacement responses presented in Figure 10 in the
paper of Bui et al. (2008) were obtained with the SPH model similar to the FE model
used by Chen and Mizuno (1975). It means that its width is actually narrower than the
current SPH model (7.32 m compared to 10 m). As pointed out by Bui et al. (2008) the
width of 7.32 m is not sufficiently wide to avoid boundary effects, especially for the
associated flow rule case;
- Effect of the artificial viscosity: as proposed by Monaghan (1992) the artificial
viscosity is proportional to the sound speed of soil. In the current SPH model, the
sound speed is smaller than the value used in the SPH model 2008 (270 m/s compared
to 600 m/s). It is known that in the SPH method, the utilization of the artificial
viscosity is necessary to stabilize the numerical solution (see Bui et al., 2008).
However, a high viscosity may result in an unrealistic stiff behaviour of the soil, and in
consequence the calculated resistance of the soil can be overestimated. One other
important advantage of using low sound speed is that the computational time of SPH

464
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

simulation can be considerably reduced since the leapfrog time stepping scheme used
in the SPH method is inversely proportional to the sound speed.
It is also obvious from Fig. 3 that both the SPH and FE methods predict almost the same
ultimate bearing capacities of 1.04 MPa for the associated flow rule and 0.96 MPa for the
non-associated flow rule. Note that they are significantly smaller than the bearing capacity of
1.21 MPa predicted by using Terzaghi’s solution (Equation (26)). These differences can be
attributed to several causes, the most important of which is that Terzaghi’s method assumes
the soil behaves as a rigid-perfectly plastic material which fails abruptly when the bearing
capacity of the soil is reached. In contrast, the Drucker-Prager constitutive model
implemented in the SPH and PLAXIS codes assumes that the soil behaves more like an
elasto-plastic material which allows the soil to deform under applied loads. Furthermore, the
soil can yield gradually and progressively due to the nature of the finite formulation, which
means that a yielding material point can cause next to it to yield until a well-defined failure
surface similar to the one shown in Fig. 2 is formed.
Fig. 4 and Fig. 5 illustrate the development of failure depicted by the shading contours of
total share strains at three different stages of footing settlement (at 5 cm, 10 cm and at 30 cm)
for the associated flow rule and the non-associated flow rule, respectively. It can be seen that
the SPH model and the PLAXIS FE model predicts more or less the same failure surfaces for
the associated and non associated flow rules. Moreover, it can be immediately noticed from
Fig. 4 and Fig. 5 the presence of three major failure zones at 30 cm settlement: a triangular
zone directly under the foundation, a Prandtl’s radial shear zone and a Rankine passive zone,
which agree very well with the general failure mechanisms assumed by Terzaghi (1943) for a
rigid strip footing.
In addition, comparison of the shadings of shear strains at 30 cm settlement of the footing
presented in Fig. 4 and Fig. 5 shows that the failure mechanisms obtained for the associated
flow rule is larger than those for the non-associated flow rule, especially in the radial shear
zone as well as in the passive zone. This is basically due to the nature of dilatancy in soil
during plastic flow. It is worthwhile to recall that for the associated flow rule case, the
dilatancy angle is set equal to the friction angle while it is equal to zero (no plastic volumetric
strain) for the case of a non-associated flow rule in all numerical simulations presented in this
paper. It is also noticed from Fig. 4 that the radial shear zone at 30 cm settlement of the
foundation predicted by SPH is much smaller than that of the FE model for the associated
flow case, although they are almost the same at 5 cm and 10 cm settlements of the
foundation. It is probably due to too large amount of dilatancy, i.e. too large plastic
volumetric strains, caused by a relatively high loading speed applied on the footing. This
effect does not influence the results obtained with the non-associated since the dilatancy
angle is set equal to zero.

465
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

1.4

Average pressure  beneath footing (MPa)
Terzaghi's analytical solution
1.2

0.8

0.6

0.4 SPH_Associated (Bui, 2008)
SPH_Associated (current)
0.2
PLAXIS_Associated
0
0 2 4 6 8 10
Vertical displacement  (cm)

1.4
Average pressure  beneath footing (MPa)

Terzaghi's analytical solution
1.2

0.8

0.6

0.4 SPH_Non‐Associated  (Bui, 2008)


SPH_Non‐Associated  (current)
0.2
PLAXIS_Non‐Associated
0
0 2 4 6 8 10
Vertical displacement  (cm)

Fig. 3. Pressure-displacement curves for both the associated flow rule (top) and non-associated flow rule
(bottom). Comparison of SPH results, FEM results and Terzaghi’s solution.

466
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

0 0.05 0.10 0.15 0.20 0.25 0 0.05 0.10 0.15 0.20 0.25

5 cm 5 cm

10 cm 10 cm

30 cm 30 cm
(a) SPH results for the associated flow rule (b) PLAXIS results for the associated flow rule

Fig. 4. Failure modes depicted by the shadings of shear strains at three levels of settlement of the footing:
5cm, 10 cm and 30 cm for the case of associated flow rule. (a) SPH results and (b) PLAXIS FE results.

0 0.05 0.10 0.15 0.20 0.25 0 0.05 0.10 0.15 0.20 0.25

5 cm 5 cm

10 cm 10 cm

30 cm 30 cm
(a) SPH results for the non‐associated flow rule (b) PLAXIS results for the non‐associated flow rule

Fig. 5. Failure modes depicted by the shadings of shear strains at three levels of settlement of the footing:
5cm, 10 cm and 30 cm for the case of non-associated flow rule. (a) SPH results and (b) PLAXIS FE results.

467
Proceedings of the 2nd International Symposium on Computational Geomechanics (COMGEO II), 2011

6 CONCLUSIONS
The application of the smoothed particle hydrodynamics (SPH) method in conjunction with
the Drucker-Prager material model in analyzing bearing capacity and failure mechanism of
shallow foundation has been presented in this paper. Fairly good agreement in the computed
results obtained from the SPH simulations and the finite element (FE) simulations using the
PLAXIS code demonstrates the accuracy as well as the powerful capabilities of the SPH
method in modelling of geotechnical problems.

7 ACKNOWLEDGEMENT
The research has been performed as a part of "GEO-INSTALL" (Modelling Installation
Effects in Geotechnical Engineering) project. The work is partially funded by the European
Community through the program "Marie Curie Industry-Academia Partnership and
Pathways", under Contract No PIAP-GA-2009-230638, and partially supported by JSPS
through Grants-in-Aid for scientific research. All these supports are gratefully acknowledged.
The first author wishes to thank the Japan Society for the Promotion of Science (JSPS) for a
JSPS Postdoctoral Fellowship.

8 REFERENCES
Bonet, J. and Lok, T.S.L. (1999), “Variational and momentum preservation aspects of smooth
particle hydrodynamic formulations”, Computer Methods in Applied Mechanics and
Engineering, 180, 97-115.
Bui, H.H., Fukagawa, R. and Sako, K. (2010), “Slope stability analysis and discontinuous
slope failure simulation by elasto-plastic smoothed particle hydrodynamics (SPH)”,
Géotechnique, in press.
Bui, H.H., Fukagawa, R., Sako, K. and Ohno, S. (2008), “Lagrangian mesh-free particles
method (SPH) for large deformation and failure flows of geomaterial using elastic-plastic
soil constitutive model”, International Journal for Numerical and Analytical Methods in
Geomechanics, 32(12), 1537-1570.
Bui, H.H., Sako, K. and Fukagawa, R. (2007), “Numerical simulation of soil-water
interaction using smoothed particle hydrodynamics (SPH) method”, Journal of
Terramechanics, 44(5), 339-346.
Chen, W.F. and Mizuno, E. (1990), Nonlinear analysis in soil mechanics. Amsterdam:
Elsevier Science Publishers.
Gingold, R.A. and Monaghan, J.J. (1977), “Smoothed particle hydrodynamics-theory and
application to non-spherical stars”, Monthly Notices of the Royal Astronomical Society,
181, 375-389.
Liu, G.R. and Liu, M.B. (2003), Smoothed particle hydrodynamics: a meshfree particle
method. World Scientific, Singapore.
Lucy, L.B. (1977), “A numerical approach to the testing of the fission hypothesis”,
Astronomical Journal, 82(12), 1013-1024.
Monaghan, J.J. (1992), “Smoothed particle hydrodynamics”, Annual Review of Astronomy
and Astrophysics, 30, 543-574.
Monaghan, J.J. and Lattanzio, J.C. (1985), “A refined particle method for astrophysical
problems”, Astronomic and Astrophysics, 149-135.
Terzaghi, K. (1943), Theoretical soil mechanics, 5th ed., John Wiley & Sons Inc., New York,
N.Y.
Zienkiewicz, O.C., Humpheson, C. and Lewis, R.W. (1975), “Associated and nonassociated
visco-plasticity and plasticity in soil mechanics”, Géotechnique, 25(4), 671–689.

468

View publication stats

You might also like