Application of MPM To Large Deformation

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

Application of MPM to large deformation analysis of embankment

Takatoshi Kiriyama1
1
Institute of Technology, Shimizu Corporation
Etchujima 3-4-17, Koto-ku, Tokyo 135-8530, JAPAN
kiriyama@shimz.co.jp

Introduction
The Material Point Method (MPM) is a means of particle-based numerical analysis that uses a
numerical grid. The numerical grid does not distort because it is reset at each time step and
each material point carries all information about the state. These characteristics make MPM
suitable for the analysis of large deformations without mesh distortion. In this study, a
derivative of MPM known as GIMPM is applied to a road embankment failure experiment, in
which the embankment is subjected to an experimental earthquake, in order to examine large
deformations.

Earthquake Response Analysis with Material Point Method


The procedure for updating grid point acceleration and velocity in MPM is described as
, , ,
(1a)
∑ , , (1b)
where k is the time step and g and p are the material quantities at the grid point and the
material point, respectively. Then, ag, vg, fgint, fgext, fgdamp, and mg are the acceleration, velocity,
internal force, external force, damping force, and mass at the grid point, respectively. Sp, mp,
vp are the shape function value, mass, and velocity at the material point and np is the number
of material points in the reference cell[1].
To simulate an earthquake response experiment, it is necessary to introduce the time history
boundary condition. This is easily realized by replacing Eq.(1a) and Eq.(1b) with the
following equations at the grid points subject to the time history boundary condition:
(2a)
(2b)
where acceleration(t) and velocity(t) are, for example, the time histories of the acceleration
and velocity of the base ground. To simulate a soil box experiment, this boundary condition is
applied to the grid points at the soil box.
In this paper, Rayleigh damping is also introduced in order to remove the high-frequency
response. The Rayleigh damping force at a grid point is described as
,
∙ ∙ ∙ ∙ (3a)
Here, the second term is the increment form of the grid point internal force. This then yields
, ,
∙ ∙ ∙∆ ⁄∆ (3b)
To verify these proposed equations, a wave propagation analysis was performed. The
numerical model used in this analysis is explained in Figure 2 and the parameters are listed in
Table 1. A Ricker wavelet (Figure 1) is input at the bottom of this analytical column in
accordance with Eq.(2).
Table 1. Parameters for wave propagation 1

acceleration (m/s )
2
simulation 0.5 (a)
Items Values
0
Particles Per Cell 1
-0.5
Dimensions(H x W) 0.125 m x 100.0 m
Width of cell 0.125 m -1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Time increment 0.00005 time (s)
0.02
 0.5
Damping factor

velocity (m/s)
0.01 (b)
 0.0005 s
Number of Particles 0
material point 800 -0.01
cell 800 -0.02
grid point 1,602 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
time (s)
Material Values
Figure 1. Time history of normalized Ricker wavelet (a)
Vs 200 m/s acceleration and (b) velocity when fp=10(Hz) and ts=0.2(s)
2
G 80000 kN/m fp : Center frequency of Fourier spectrum
 2.0 g/cm
3
ts : Time of maximum amplitude

unit : (m)
0.125 x 800 = 100

. . . .
grid point
bottom material point surface
0.125
Figure 2. Numerical model for wave propagation simulation. The Ricker wavelet is input at the bottom
grid point and then propagates toward the ground surface where it repeatedly reflect within the ground

No damping Rayleigh damping =0.5, =0.0005


2 2
(a) depth = 0.1 m (MPM), 0.0 m (FEM) (d) depth = 0.1 m (MPM), 0.0 m (FEM)
acceleration
acceleration

1 1
(m/s )

(m/s )
2
2

0 0
-1 -1
-2 -2
2 2
(b) depth = 49.9 m (MPM), 50.0 m (FEM) (e) depth = 49.9 m (MPM), 50.0 m (FEM)
acceleration

acceleration

1 1
(m/s )

(m/s )
2

0 0
-1 -1
-2 -2
2 2
(c) depth = 99.9 m (MPM), 99.9 m (FEM) MPM (f) depth = 99.9 m (MPM), 99.9 m (FEM) MPM
acceleration
acceleration

1 1
FEM FEM
(m/s )

(m/s )
2

0 0
-1 -1
-2 -2
0 0.5 1 1.5 2 0 0.5 1 1.5 2
time (s) time (s)
Figure 3. Acceleration time histories obtained from wave propagation simulation. Figures (a), (b), and (c) are the
results at the surface, center, and bottom of the column with no damping. Figures (d), (e), and (f) are the results
at the same depths with Rayleigh damping

Figure 3 shows the time histories resulting from the proposed procedure. The waveforms at
the center and the bottom of the column (Figure 3(b) and (c)) have the same amplitude as the
input wave, while the amplitude is double at the top of the column under the no-damping
condition (Figure 3(a)). After the wave is reflected at the bottom, the amplitude inverts and
the propagation continues as before. This result agrees with the wave equation, verifying
Eq.(2a) and Eq.(2b). The waveform obtained when Rayleigh damping is applied is similar to
that under no-damping condition; only the amplitudes are reduced because of Rayleigh
damping force. To demonstrate the effectiveness of this formulation of the Rayleigh damping
force, conventional FEM results are shown in Figure 3 for reference. The time histories of
acceleration given by MPM and FEM are the same, verifying that Eq.(3a) and Eq.(3b)
function correctly as a Rayleigh damping force.

Particles Per Cell Investigation


The number of Particles Per Cell (PPC) has been discussed in relation to mesh-based particle
methods such as PIC and MPM, and it has been reported that more than two particles are
necessary on each axis for an accurate simulation in continua. On the other hand, there is a
need to investigate PPC for discrete materials, such as soil, because slope failure exhibits
intensively discrete behavior.
Figure 4 shows the final deformations of sand columns after analytically removing a rigid
retaining wall. Here the sand material is as specified in Table 2. The spread of sand in Figure
4(c) and Figure 4(d) is apparently insufficient because the contact logic of MPM is controlled
not by the material point but by the cell size. That is, multiple material points in one cell
induce an excess dilatancy in discrete materials. Following this result, PPC=1 is adopted in
this paper.
Table 2. Analysis parameters for 0.25
vertical position (m)

sand spreading simulation 0.2 (a) PPC=1 (c) PPC=4


Items Values 0.15 Particles
Particles Per Cell 1, 4, 9 0.1 Initial Shape
Dimensions(H x W) 0.25 m x 1.0 m 0.05 Rigid wall
Width of cell 0.005 m
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
Time increment 0.00005 s horizontal position (m) horizontal position (m)
Damping factor 0.0 0.25
vertical position (m)

Number of Particles 0.2 grid point (b) PPC=1 (d) PPC=9


material point 1600, 6400, 14400 0.15 cell
material point
cell 10,000 0.1
grid point 10,251 0.05
Material Values 0
2
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
G 0i 7280 kN/m horizontal position (m) horizontal position (m)
3
 1.6 g/cm Figure 4. Comparison of sand column spread after removing rigid wall with
 40 deg different PPC (particles per cell). (a) Initial particle position; (b), (c), and
2
c 0.0 kN/m (d) Sand spread with PPC=1, 4 and 9

Comparison Between Experiment and Simulation


An experiment was carried out on a road embankment consisting of DL Clay as the soil
material. The embankment is subjected to a centrifugal force of 30g. To simulate an
earthquake, a horizontal acceleration was applied to the soil box to induce failure of the
embankment. Surface deformation of the embankment, as well as the shear band that
appeared within the embankment, was observed. Figure 5 shows the placement of
instrumentation[2]. The same road embankment was then numerically simulated as a
collection of particles. Each particle was given a geotechnical nonlinearity in the form of a
Mohr-Coulomb constitutive model. The parameters for this simulation are listed in Table 3.
After performing a static gravitational analysis to obtain the initial stress, the earthquake
waveform was applied to the grid point at the soil box.
Figure 7 and Figure 8 compare the experimental and simulated deformations. Settlement is
slightly greater in the simulation than in the experiment, while spreading is overestimated in
the simulation. The simulation shows settlement around the crown of the embankment (Figure
7(b)), which is not seen in the experiment. However, both simulation and experiment clearly
indicate a shear band localization. As a whole, the shape of the simulated post-earthquake
embankment is in good agreement with the deformation observed in the experiment.
900 5

acceleration (m/s2)
[27000]
335 350 200 2.5 (a)
[10050] [10500] [6000] 15
DV3 100 [450]
DH3 EPS 0
Laser Displacement Sensor 60 -2.5
1:1.2 DH2 AH3 AV3
-5
Bender Element 300
0 5 10 15 20 25
Accelerometer
[9000]
0.5 time (s)
292
Strain Gauge

velocity (m/s)
DH1
100
[8760]
0.25 (b)
Colored Sand 232
146
Laser Displacement Sensor
AH2 AV2 [6960] 0
90
-0.25
Origin (0,0) -0.5
AH1 AV1
Silica No.3 Base Sheet unit : (mm) 0 5 10 15 20 25
time (s)
Figure 5. Instrumentation overview. Figures in square brackets are
converted ones multiplied by the 30g centrifuge magnification[2] Figure 6. (a) Observed acceleration and
(b) calculated velocity, which are the
earthquake inputs for the simulation
Table 3. Analysis parameters for road embankment simulation
Items Values Number of Particles Material Values
2
Particles Per Cell 1 material point 6,316 G 0i 160 kN/m
3
Dimensions(H x W) 10.0 m x 25.0 m cell 16,000  1.53 g/cm
Width of cell 0.125 m grid point 16,281  30.5 deg
Time increment 2
0.00025 s c 8.5 kN/m
 0.0
Damping factor
 0.001

Top
(a) (b)
unit : (m)
initial position 2.7
after shaking Center

Bottom 2.6
9.0

2.4

Figure 7. (a) Observed deformation and shear strain localization inside embankment after an earthquake[2];
(b) Maximum shear strain contour obtained from MPM simulation

Bottom Center Top


5 5 5
4 (a) (c) (e)
settlement (m)

4
settlement (m)

4
settlement (m)

3 Simulation 3 3
2 2 2
1 1 1
0 Observed 0 0
-1 -1 -1
5 5 5
4 (b) 4 (d) 4 (f)
spreading (m)

spreading (m)

spreading (m)

3 3 3
2 2 2
1 1 1
0 0 0
-1 -1 -1
0 5 10 15 20 25 0 5 10 15 20 25 0 5 10 15 20 25
time (s) time (s) time (s)
Figure 8. Settlement and lateral spread time histories of targets on slope surface
References
[1] T. Kiriyama: Numerical simulations of progressive failure of triaxial compression tests
using generalized interpolation material point method, Journal of Applied Mechanics, JSCE,
Vol.16, 2013. (In Japanese)
[2] K. Tokida, J. Jang, K. Oda, A. Nakahira and A. Ohtsuki: Centrifuge test on advanced
seismic performance of road embankment against sliding failure during earthquake, Journal
of Earthquake Engineering, JSCE, Vol.29, pp.637-645, 2007. (In Japanese)

You might also like