Sie sind auf Seite 1von 15

A corrected smooth particle hydrodynamics formulation

of the shallow-water equations


Miguel Rodriguez-Paz, Javier Bonet
*
Civil and Computational Engineering Centre, School of Engineering University of Wales Swansea, Swansea SA2 8PP, UK
Accepted 30 November 2004
Abstract
A shallow-waters formulation based on a variable smoothing length SPH method is presented. This new formulation
of the SPH equations treats the continuum as a Hamiltonian system of particles where the constitutive relationships for
the materials are introduced via an internal energy term. Some of the advantages of the new SPH formulation are evi-
dent in the solution of the shallow-water equations for expanding ows. The shallow-waters approach incorporates the
terrain into the equation of motion through terrain properties evaluated using SPH methodology. Several examples are
presented on the simulation of breaking dams on dierent geometries. A comparison with the analytical solutions is also
included.
2005 Elsevier Ltd. All rights reserved.
Keywords: Shallow-waters; SPH; Breaking dam; Free surface ows
1. Introduction
During the last decade, a number of changes in glo-
bal climate and other man-induced changes in nature
such as deforestation and pollution, have triggered envi-
ronmental problems, in particular water related issues:
oods and mudslides. On the other hand, the manage-
ment of water resources is also a main area of research,
which includes the prediction of storm surges, hazard
prediction of dam breaks, sediment transport and coast-
al tides. Flow models that realistically represent the
physical properties of the ow and the complex topo-
graphic that are found in regions where debris ava-
lanches occurrence is high, can help in the hazard
prediction of such phenomena and help to mitigate their
destructive power. Numerical techniques are a viable
alternative when the phenomenon is dicult to repro-
duce in the laboratory due to many factors, such as,
its scale and magnitude.
Many of the hydrodynamic models for reservoirs and
tidal predictions are based on the solution of the depth
averaged shallow-water equations using nite dierences
or nite element procedures. Other numerical methods
used for the solution of the shallow-water equations
for bore wave propagation include the nite element
method and more recently nite volume methods [14].
Most of these methods are based on elements or cells,
with the dependence on the grid and mesh renement
to resolve the complex topography and evolving ow
features. However, most of the techniques that have
0045-7949/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compstruc.2004.11.025
*
Corresponding author. Tel.: +44 1797 295 689; fax: +44
1792 295 676.
E-mail addresses: rodriguez.miguel@itesm.mx (M. Rodri-
guez-Paz), j.bonet@swan.ac.uk (J. Bonet).
Computers and Structures 83 (2005) 13961410
www.elsevier.com/locate/compstruc
been developed based on the shallow-water equations,
assume a small gradient of the terrain and do not con-
sider a vertical component of the velocity.
Meshless methods, and in particular Smooth Particle
Hydrodynamic (SPH) methods have already been used
to simulate free surface ows [58]. SPH is a robust
numerical technique introduced by Lucy and Gingold
[9,10]. Since its introduction in astrophysics the method
has been applied to simulate problems with complicated
physics such as multiphase ows [11] and high strain
dynamics. SPH has also been applied to simulate geo-
physical ows like debris ows [12] and ice elds [13].
In this paper, a novel variational formulation of the
Lagrangian shallow-water equations is presented. This
formulation uses the variable smoothing length ap-
proach for SPH developed by Bonet [14] and which will
be also presented in the following sections. This new
methodology is intended to deal with problems of ows
over a steep and non-uniform general terrain, like in the
case of avalanches and debris ows.
As the algorithm presented is explicit and solves only
two components of the 3-D space, the storage require-
ments are minimum and can be implemented on per-
sonal computers or small workstations. This fact
presents some importance to practising engineers and
geophysicists who may wish to experiment with the
code.
2. General assumptions
The shallow-water assumption is based on a 2-D plan
view projection of the problem domain. In this way, for
the case of a SPH discretisation of the resulting 2-D do-
main, each particle represents a column of uid of a cer-
tain height. These particles move according to the
topography of the terrain but always in a direction tan-
gent to the terrain.
Consider a plan view of a terrain as shown in Fig. 1.
The terrain is represented by a general function H(x, y),
which gives the height at each point. The continuum is
discretised with a system of Lagrangian particles, in
which each particle represents a column of water of
height h
t
with constant mass m, which moves over the
terrain. The basic assumption is that the velocity
through all the height of the vertical column is uniform
and parallel to the terrain. This implies that the instan-
taneous spatial variation of h
t
is small. The motion of
the Lagrangian particles is then followed in time.
The motion of the columns of water is constrained to
follow the terrain. This implies that the global z position
of the bottom of each column is given by (see Fig. 2)
z Hx; y 1
Dierentiating with respect to time, the vertical compo-
nent of the velocity can be evaluated as
v
z
rH v 2
where $H is the gradient of the terrain at the current po-
sition occupied by the column and v = (v
x
, v
y
) is the 2-D
velocity vector containing the x and y velocity compo-
nents of the column. The unknowns of the problem
are therefore the x and y co-ordinates of every particle
at each time and the height of the water column h
t
.
Alternatively, instead of the column heights h
t
, it is also
possible to use a related variable given by the 2-D pro-
jected density of the uid q, that is the amount of mass
per unit of area. Given that the uid in motion will be
assumed to be incompressible, this density and the col-
umn height are simply related by
q h
t
q
w
3
where q
w
denotes the constant 3-D density of the uid.
Using this variable instead of the column height renders
the problem formally analogous to a standard 2-D SPH
formulation of a compressible uid.
In order to derive the governing equations, a
variational approach is used [14]. This is based on the
Fig. 1. Discretisation of the uid with a system of Lagrangian
particles that move on top of a general terrain.
Fig. 2. Velocity components of each column of water.
M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410 1397
expression of the total energy of the system as a function
of the particle positions and will therefore require the
evaluation of the projected density in terms of such par-
ticle positions.
3. Density evaluation
3.1. Standard SPH
In a standard SPH formulation with constant
smoothing length, the density at a particle is smoothed
over a sub-domain that is dened by a circle of radius
2h, where h is the smoothing length of a kernel function
W (see Fig. 3) [6].
If the particles have a mass m
I
, the discrete system of
particles has a density function of the form
^ qx

I
m
I
dx x
I
4
where d is the Dirac delta. In order to reconstruct the
smooth continuum, the kernel-smoothing concept can
be applied to the discrete density (4) to give a continuum
density approximation over a domain D as
qx
_
D
^ qx
0
W
x
x
0
dx
0
_
D
W
x
x
0
dx
0
5
where W
x
0 x; h W x; x
0
; h represents a suitable kernel
function [16]. Substituting Eq. (4) into the above expres-
sion and noting the Diracs delta properties, gives
q
I

J
m
J
W
I
x
J
; h
I

_
D
W
I
x
J
; h
I
dV
6
Typically, the kernel function is scaled so that the inte-
gral in the denominator becomes one. However, in order
to deal with rigid boundaries, Bonet et al. [14,15] have
introduced a function gamma as
c
I
x
I
; h
I

_
D
W
I
x; h
I
dV 7
and substituting it into Eq. (6) gives
q
I

J
m
J
W
I
x
J
; h
I

c
I
x
I
; h
I

8
In the standard SPH formulation and in the absence of
boundaries the gamma function is c
I
(x
I
, h
I
) = 1. This is
no longer valid for particles that are within a distance
to a rigid boundary that is less than 2h. The evaluation
of the integral dened by Eq. (7) can be complicated for
the case in which a particle is within certain distance
from a rigid boundary. Kulasegaram et al. [15] propose
a numerical method to approximate this integral. In the
following sections, however, we assume there are no ri-
gid boundaries, and therefore c = 1. This is appropriate
for many shallow-water applications.
3.2. Variable-h SPH
The evaluation of the density using a constant
smoothing length is normal practice in SPH. In the case
of uids with certain compressibility, however, a simple
case of uniform expansion or contraction can only be ex-
actly represented if the smoothing length is allowed to
vary. In the shallow-water approximation, the uid will
follow the terrain and its projected 2-D density will ex-
pand or contract according to the height of the water
column as shown by Eq. (3). A variable smoothing
length is therefore needed in order to maintain accuracy
of the solution. In general, h must change according to
[16]
qh
d
m
constant q
0
h
dm
0
9
where d
m
is the number of space dimensions, q
0
, h
0
are
the initial density and smoothing length, respectively,
and q and h are the current values of density and
smoothing length for any particle. This gives an equa-
tion for the instantaneous smoothing length h for a par-
ticle I as
h
I
h
0
q
0
q
I
_ _
1=dm
10
It is important to note that the above equation for the
density is implicit as the density q
I
is itself a function
of h
I
. In particular, the density is evaluated according
to Eq. (6), which for the case where there are no rigid
boundaries (i.e., c = 1) gives
q
I

J
m
J
W
I
x
J
; h
I
11
Note that this is now a non-linear equation for q
I
due
to the above dependency of h
I
on q
I
. Fortunately, the
equation for each particle are not coupled and can there-
fore be solved independently. A simple NewtonRaph-
son iteration to achieve this is described in Section 5.1
below. Fig. 3. Particle interpolation and kernel function.
1398 M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410
4. Equations of motion: non-dissipative case
Consider now the uid described by a system of SPH
particles that are located in a 2-D Cartesian space (x, y),
the position of each particle dened by the vector x
I
and
the horizontal component of the velocity of the particle
by v
I
. Each particle I will represent a column of water
with a total mass m
I
, which will remain constant during
the motion and hence, conservation of mass will be
ensured.
The EulerLagrange equations of motion of the sys-
tem of particles are, in the absent of dissipative forces,
e.g., bottom friction and viscosity eects [17]:
d
dt
oL
ov
I

oL
ox
I
0 12
where the Lagrangian functional L is dened in terms of
the kinetic energy K and potential energy p, as
L K p 13
and noting that p is only a function of the positions of
particles x
I
, Eq. (12) can be written as
d
dt
oK
ov
I

oK
ox
I

op
ox
I
14
The expressions for the kinetic and potential energies are
presented in the following sections.
4.1. Kinetic energy
The kinetic energy of the system of particles can be
approximated as the sum of the kinetic energy of each
particle, which with the help of Eq. (2), gives
K
1
2

I
m
I
v
I
v
I
v
2
z
; v
z
v
I
rH
I
15
It is important to notice the term v
2
z
v
I
rH
I

2
, which
takes into account the vertical component of the veloc-
ity. This term is often neglected in other shallow-water
approaches, since it is considered to be too small for
small terrain gradients. In order to be able to deal with
large gradients in the terrain, it is important to retain it
throughout the derivation of the equations of motion.
4.2. Potential energy
The potential energy of each particle is calculated at
the centre of gravity of each water column, i.e., H
1
2
h
t
.
Hence, the total potential energy of the system of parti-
cles can be expressed as the sum of the potential energy
of each particle
p

I
m
I
gH
I

1
2

1
m
I
gh
t;I
16
where g denotes the gravity acceleration. The rst
term represents the external energy term, whilst the sec-
ond term can be interpreted as the pseudo-internal
energy,
p
ext

I
m
I
gH
I
17
p
int

I
m
I
1
2
gh
t;I
_ _
18
The internal energy component can be expressed in
terms of an internal energy per unit mass w(q) as
p
int

I
m
I
wq
I
19
where
wq
I

1
2
g
q
I
q
w
_ _
20
This expression matches that presented in [14] for gen-
eral uid applications. The corresponding pressure term
dened as p q
2 dw
dq
becomes the well-known height inte-
grated hydrostatic pressure, as usual in the case of shal-
low-water applications, and is given by
p q
2
dw
dq

1
2
gq
w
h
2
t
21
4.3. Evaluation of inertial forces
Eq. (14) can be re-written in the more usual format of
Newtons second law as
d
dt
oK
ov
I

oK
ox
I
..
inertial forces
F
I
T
I
; F
I

op
ext
ox
I
; T
I

op
int
ox
I
22
where the left-hand side represents the inertial forces of
the system, F
I
and T
I
are the external and internal forces
of the system, respectively, which will be discussed in de-
tail in the section below.
The equation of motion can also be expressed as
I
I
F
I
T
I
23
where the inertial forces I
I
can be evaluated with the help
equation (15) as
I
I

d
dt
oK
ov
I

oK
ox
I
24
and substituting the expression for K gives after simple
algebra:
I
I

d
dt
m
I
v
I
m
I
v
I
rH
I
rH
I
m
I
v
I
rH
I
k
I
v
I
25
where k
I
is the curvature tensor of the surface H(x, y) at
the position occupied by particle I and is given by
k rrH 26
M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410 1399
Note that the last term in Eq. (25) could be interpreted
as the centripetal acceleration of the column of water.
4.4. External forces
The external forces can be evaluated with the help of
Eqs. (22) and (17) to give
F
I

op
ext
ox
I
m
I
grH
I
27
4.5. Internal forces
The internal constitutive forces are calculated using
the procedure described in [13] and are given by the
expression
T
I

op
int
ox
I
28
The evaluation of this term, in general, requires the con-
stitutive denition of the material. For the case of uids
with no viscosity, the internal forces are calculated using
Eqs. (19) and (11) as
T
I

op
int
ox
I

J
m
I
m
J
p
J
a
J
q
2
J
rW
J
x
I
; h
J

p
I
a
I
q
2
I
rW
I
x
J
; h
I

_ _
29
where a is a correction factor that emerges in the vari-
able-h formulation (for details see [14]):
a
I

J
m
J
r
IJ
dW
I
dr
IJ
30
and p denotes the hydrostatic pressure given by Eq. (21).
4.6. Acceleration
An expression for the acceleration of the particles can
be found by substituting Eqs. (25), (27) and (29) into Eq.
(23) and solving for a
I
gives, after some simple algebra:
a
I

g v
I
k
I
v
I
t
I
rH
I
1 rH
I
rH
I
rH
I
t
I
31
where t
I
= T
I
/m
I
. Eq. (31) includes a term for the curva-
ture of the terrain as well as one term for the gradient of
the terrain. It can be noted that for the case of a at ter-
rain ($H = 0; k = 0), Eq. (31) becomes a
I
= t
I
. In gen-
eral, Eq. (31) can be easily implemented in an explicit
time integration scheme. For the examples included in
this paper, the values of $H and k
I
can be easily evalu-
ated nding the derivatives of the equation that denes
the surface for the terrain. For more general applica-
tions, an interpolation of the terrain from given grid
points will be necessary.
5. Numerical implementation
5.1. Numerical evaluation of density
As previously mentioned, Eq. (11) is non-linear in q
I
.
In order to solve it, a NewtonRaphson solution proce-
dure can be used. Let us dene a residual Rq
k
I
for each
iterative value k of the density as
Rq
k
I
q
k
I

J
m
J
W
I
x
J
; h
k
I
32
where the smoothing length h is a function of the density
through Eq. (10) as
h
k
I
h
0
I
q
0
I
q
k
I
_ _
1=d
m
33
and h
0
I
and q
0
I
are the initial values of h
I
and q
I
,
respectively.
A simple possible iterative solution for q
I
is given by
q
k1
I

J
m
J
W
I
x
J
; h
I
q
k
I
34
However, a much better option is to use a Newton
Raphson solution by the iteration of
q
k1
I
q
k
I

R
k
I
dR
dq
_ _
k
I
35
Substituting Eq. (32) gives the iterative solution as
q
k1
I
q
k
I

q
k
I

J
m
J
W
I
x
J
; h
k
I

dR
dq
_ _
k
I
36
where the derivative of the residual is calculated as
dR
dq
1

J
m
J
dW
I
dh
I
dh
I
dq
I
1

J
m
J
q
I
d
m
W
I
d
m
r
IJ
dW
I
dr
IJ
_ _
37
After simple algebra, this last equation becomes
dR
dq
1
1
q
I

J
m
J
W
I
x
J
; h
I

a
I
d
m
q
I
38
Finally, substituting Eq. (38) into Eq. (35), gives
q
k1
I
q
k
I
1
R
k
I
d
m
R
k
I
d
m
a
k
I

_ _
39
which is the NewtonRaphson approximate solution for
the density equation at iteration k + 1.
In order to start up the iteration process an initial
guess is required. For this purpose, the equation for
the rate of change of density [18] can be integrated in
time to give
1400 M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410
_ q
I

q
I
d
m
a
I

J
m
J
rW
I
x
J
; h
I
v
J
v
I
40
which enables a simple guess of the density at step n + 1
to be evaluated in terms of the density at step n as
q
0
I;n1
q
I;n
e
k
n
41
where
k
d
m
Dt
a
I

J
m
J
v
J
v
I
rW
I
x
J
; h
I

_ _
42
5.1.1. Convergence
In order to stop the NewtonRaphson iterative pro-
cedure, a tolerance that is within the machine precision
must be dened. Convergence is achieved when
jR
k1
I
j
q
k
I
6 e 43
Typical values in a double precision machine are
e = 10
15
, which is usually achieved within a few itera-
tions. A cheaper alternative is to evaluate the density
using Eq. (41) and the value of the smoothing length
from the previous time step.
Once the new values for the density have been evalu-
ated, new heights of the water columns can be obtained
using Eq. (3) or
h
t;I

q
I
q
w
44
It is important to mention that by evaluating the density,
new values for the smoothing length and the correction
factor a are also calculated in each iteration. The up-
dated values of the smoothing lengths are then used in
all the subsequent SPH interpolations within the time
step.
5.2. Time integration scheme
The equation of motion is assembled using Eqs. (23)
(31) and in order to update the position of particles, an
explicit time integration scheme is used, namely the leap-
frog method, dened as
v
n1=2
I
v
n1=2
I
Dta
n
I
45
x
n1
I
x
n
I
Dt
n1
v
n1=2
I
46
where
Dt
1
2
Dt
n
Dt
n1
47
Due to the explicit nature of the scheme, the Courant
FriedrichsLewy (CFL) stability criteria must be
satised. This implies that the time step size must be less
than
Dt CFL
h
min
maxc
I
kv
I
k
; 0 6 CFL 6 1.0 48
where c is the wave speed of propagation or speed, de-
ned as [19]
c
I

gh
t;I
_
49
and h
min
is the minimum smoothing length of the system
of particles. The magnitude of the velocity is also consid-
ered. Although this equation should provide time steps
that would satisfy the stability condition, in the numer-
ical examples presented in the following section, the
CFL factor considered was <10% in all the cases. Fur-
ther analysis is needed to nd an upper limit for the time
step that could guarantee accurate results without the
loss of accuracy.
6. Numerical examples
6.1. Introduction
This section presents several examples for the appli-
cation of the method with breaking-dam problems on
dierent terrains. In all the cases the uid considered is
water. No viscosity eects are considered in the uid
model but they can be included as part of a dissipative
potential, aecting the internal forces.
6.2. Breaking dam in rectangular channel
An example for the validation of the code, for which
an analytical solution exists, is the case of a innite
breaking-dam. A dam of certain depth and innite
transversal dimension brakes at time t = 0 s. This prob-
lem is clearly of interest in a number of real situations,
ranging from the catastrophic ood following the col-
lapse of a dam, to the operation of sluice gates in an irri-
gation channel.
An analytical solution exits for the problem and is
presented in Ref. [20]. The results provided by the SW-
SPH method are compared to those obtained with the
analytical solution.
6.3. Problem set-up
For the numerical simulation, only a strip of 1 m was
considered, as shown in Fig. 4. On the right-hand side of
the dam, at x = 2 m there is a gate that is instanta-
neously removed at time t = 0.0 s.
There is no slope in the channel and the friction is not
considered in order to compare the results with the ana-
lytical solution. A total of 6601 particles were used, sim-
ulating the solid wall on the left with a symmetry
condition in the uid, i.e., two gates were considered:
M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410 1401
Fig. 4. Dimensions of the channel and the dam.
Fig. 5. SW-SPH results for t = 0.01 s and t = 0.10 s. Colours indicate depths. (For interpretation of colour in this gure legend the
reader is referred to the web version of this article.)
0.00
0.10
0.20
0.30
0.40
0.50
0.60
0.70
0.80
0.90
1.00
0.05 0.15 0.25 0.35 0.45 0.55 0.65
d
e
p
t
h

h
h @ x=2.0m
Analytical Solution
Depth at gate
t (s)
Fig. 6. SW-SPH results vs. analytical solution for depth at x = 2.0 m.
0.00
0.10
0.20
0.30
0.40
0.50
0.60
0.70
0.80
0.90
1.00
0.05 0.15 0.25 0.35 0.45 0.65
(
g
*
h
)
^
0
.
5
Vel @ x=2.0m
Analytical Solution
Velocity at gate
t (s)
0.55
Fig. 7. SW-SPH results vs. analytical solution for velocity at x = 2.0 m.
1402 M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410
one at x = 2 m and the other one and x = 2 m,
although the results shown in here are for the right side
(x > 0.0). For this simulation, the particles were allowed
to move only in the x direction, since it was assumed
that the dimensions of the channel are big enough to
consider only a central strip of the uid (Figs. 5 and 6).
According to the analytical solution, the depth of the
water and the original position of the gate should remain
constant and equal to 4/9 h
0
until the point where the
wave that travels backwards reaches the solid wall at
x = 0.0 m. This occurs approximately at t = 0.65 s. In this
case, the depth of the uid should be h = 0.444 m, until
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
t (s)
x

p
o
s
front position
Analytical Solution
Front position
Fig. 8. SW-SPH results vs. analytical solution for a particle near the front.
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0.00 1.00 2.00 3.00 4.00 5.00 6.00
x pos
Analytical Solution
SW-SPH
t=0.10 s
d
e
p
t
h

(
m
)
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0.00 1.00 2.00 3.00 4.00 5.00 6.00
Analytical Solution
SW-SPH
x pos
t=0.20 s
d
e
p
t
h

(
m
)
Fig. 9. Comparison of the prole for t = 0.10 s and t = 0.20 s. The dots represent the SW-SPH solution and the continuous line the
analytical solution.
M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410 1403
t = 0.64 s. The SW-SPH results are shown in Fig. 7, com-
pared to the value predicted by the analytical solution. In
the same manner, the velocity of the uid at the point of
the gate (x = 0.0) should remain constant and equal to 2/3
c
0
, where c
0

gh
0
_
3.1314 m s
1
. The results are
shown in Fig. 7, normalized with respect to c
0
.
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0.00 1.00 2.00 3.00 4.00 5.00 6.00
Analytical Solution
SW-SPH
t=0.40 s
x pos
d
e
p
t
h

(
m
)
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0.00 1.00 2.00 3.00 4.00 5.00 6.00
Analytical Solution
d
e
p
t
h

(
m
)
x pos
SW-SPH
t=0.60 s
Fig. 10. Comparison of the prole for t = 0.40 s and t = 0.60 s. The dots represent the SW-SPH solution and the continuous line the
analytical solution.
Fig. 11. Lateral view of the collapse of a cylindrical column of water. From left to right: t = 0.0 s, t = 0.10 s, t = 0.20 s and t = 0.30 s.
1404 M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410
Using the formula given in Ref. [20], the position of a
particle at the front is given by
x 3

gh
_
v
0
2

gh
0
_
t 50
If h is known for a particle located towards the front, the
position of the particle given by the program is com-
pared with the one provided by Eq. (50). Fig. 8 shows
the results for the front of the ow. The numerical re-
sults match very well those given by the analytical solu-
tion (Figs. 9 and 10).
6.4. Cylindrical dam on a horizontal plane
In this example a cylindrical column of water col-
lapses on a horizontal plane. The column has 1 m in
diameter and 1 m in height. In this case the properties
of the terrain are k = 0, $H = 0 (at terrain). The results
are shown in Figs. 11 and 12. It can be seen that the ini-
tial circular conguration is perfectly kept throughout
time, which would not be the case for a constant
smoothing length approach, as it was pointed out in
the previous chapter. A total of 5133 particles were used.
Fig. 12. Top view of the breaking dam.
M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410 1405
The results are displayed as a surface, using standard
matlab graphics.
6.4.1. Constant-h results comparison
The same example was solved using a constant-h cor-
rected SPH formulation of the shallow-water equations.
As shown in Fig. 13, the results of the constant-h ap-
proach are not able to simulate the problem, as the par-
ticles near the expanding boundary of the circle have less
neighbour particles as the simulation continues. On the
other hand, the results for the SW-SPH approach pre-
sented in this paper show a uniform distribution of the
particles, without losing the circular shape. The same
distribution of particles was used in both cases and no
bottom friction was included.
In this example the terrain was considered as per-
fectly smooth, i.e., no bottom friction.
6.5. Dam in triangular channel
In this example a channel with triangular cross-sec-
tion is used to represent the terrain on which the ow
will move. The geometric properties and dimensions
for the initial set-up are shown in Fig. 14. The problem
consists of the instantaneous breaking of a dam curtain,
which makes the whole water volume move down the
steep hill. The sequence is indicated in Fig. 17.
For this case the gradient and the curvature of the
terrain are dened as
rH
0.4
0.4 signy
_ _
; k 0
The other parameters employed for this simulation are-
g = 9.806 m s
2
; q
0
= 1000 kg m
3
; no viscosity and bot-
tom friction were considered.
A total of 3793 particles were used in the simulation.
The results are shown in Fig. 15. The initial congura-
tion is a dam that contains the water in a triangular
channel. At time t = 0.0 s the curtain is removed instan-
taneously and the water starts moving down the chan-
nel. A top view is presented for better clarity and the
colours of the graphs represent the depth of the uid
over the channel.
Once again, it is important to mention that the mesh-
less nature of the methodology allows for large changes
in the geometry of the domain of the uid to take
place without the need of re-meshing. This is also possi-
ble due to the variational consistent equations for the
density, which in a way work as an adaptive procedure
by changing the smoothing length for each particle as
changes in particle distribution occur throughout the
simulation.
6.6. Steep parabolic channel
A more complex geometry is used in this case to
model the surface on which the uid moves. It consists
of a surface dened by
z
y
2
2R
mx
where R is a radius for the parable and m is a slope in the
x-direction. In this simulation R = 1.1 m and m = 40.
The geometric properties, gradient and curvature of
the terrain are therefore
Fig. 13. (a) Standard SPH for a collapsing column of water, (b) variational SPH.
Fig. 14. Cross-section for the channel.
1406 M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410
rH
0.839
0.909y
_ _
; k
0.0 0.0
0.0 0.909
_ _
And the material parameters are g = 9.806 m s
2
; q
0
=
1000 kg m
3
. In Fig. 16, the results of the simulation
are shown in top views, i.e., XY plane.
Fig. 15. Breaking dam in a triangular channel, the colours indicate the depth of the uid over the channel. (For interpretation of
colour in this gure legend the reader is referred to the web version of this article.)
M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410 1407
The terrain is represented as contour curves. The
cylindrical column of water is released at time t = 0.0 s.
This column of water has a diameter of 60 cm and an
initial constant coordinate for the free surface of
Z = 0.10 m. A total of 4595 particles were used in this
simulation. The colours of the graphs represent in this
case the depth of the ow over the channel.
The results are for times t = 0.0 s, t = 0.50 s, t = 2.0 s
and t = 4.0 s, respectively. It is important to notice that
there are some parts of the uid which have almost nil
depth, indicate by the darker blue area, representing
wetted regions of the channel. The simulation was car-
ried out on a PC with 256 MB of physical RAM with 2
Pentium III processors with a clock speed of 550 MHz.
The total time for the simulation up to t = 4 s was
achieved in just a few hours.
6.6.1. Eccentric case
Considering now the same geometric properties of
the terrain used in the last example but now the column
of water is initially placed in an eccentric manner on the
channel. The purpose of this is to test the ability of the
method to deal with the curvature of the terrain. The
column of water was supposed to be a cylinder of
0.50 m of diameter, 0.25 m of height with centre at
x = 0.25, y = 0.50.
Fig. 17 shows the results of the simulation. It is clear
how the uid moves along the channel in a direction dri-
ven by its geometry. A sudden break initiates the ow
and then the uid nds the centre of the channel and
starts moving down the slope along it. A total of 4100
particles were used to represent the uid. The simulation
was run up to a time t = 8.0 s.
Taking this example as reference, if the same spacing
used for the initial distribution of particles used for the
uid were to be used to mesh the entire channel, the
number of particles would be approximately over
187,000. The approach followed by other numerical
techniques is based on gridding the whole terrain and
then track the ow, with wet and dry cells. This SW-
SPH approach, however, discretises only the uid and
the interaction of the terrain friction and any other dis-
sipative eect are included as forces aecting the motion
of each particle, acting at the position of that particle.
Fig. 16. Breaking dam in a parabolic channel. Dam is symmetrically located along the centre of the channel. The colours show the
depths of the ow for dierent times. (For interpretation of colour in this gure legend the reader is referred to the web version of this
article.)
1408 M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410
7. Concluding remarks
A new methodology for the numerical simulation of
shallow-water-like uids over a general terrain was pre-
sented. The new set of equations is based on the varia-
tional SPH formulation presented by Bonet et al. [14].
The formulation has been denominated: variational
SPH formulation of Lagrangian shallow-water equa-
tions. Although the new formulation includes some of
the standard shallow-water equations assumptions, it
incorporates a new treatment of the terrain, as it allows
more general terrains to be considered. The resulting
technique shows a lot of potentiality for problems deal-
ing with breaking dams, ooding, debris ows [18,21],
avalanches and tidal waves, among others. Numerical
results show good agreement with analytical solutions.
Fig. 17. Breaking dam eccentrically in a parabolic channel. A total of 4100 particles were used in the simulation. The graphs show the
depth of the ow at dierent times. Initial depth 0.25 m.
M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410 1409
The results show that the technique is robust and stable,
which is in agreement with the previous work showing
that this type of particle methods are stable in the pres-
ence of compressive pressure values [22]. Given that the
pressure in the shallow-water model is proportional to
the height squared, its compressive nature is always
physically and numerically assured. Another advantage
of the method over traditional numerical techniques that
use grids is that there is no need for a mesh for the entire
terrain; only the uid is discretised with particles. This
enables the method to be implemented in serial machines
such as personal computers, since the memory require-
ments are vastly reduced. At present the cost of the
implementation appears high. This is very possibly due
to the type of searching technique used to determine par-
ticle neighbours, which is based on the Alternating Dig-
ital Tree method [23]. It is likely that other types of
searching methodologies may be more ecient for this
type of application and should be explored in the future.
The method can also incorporate more sophisticated
constitutive models for dierent materials and bottom
friction forces. The method shows the potential to be ap-
plied in the simulation of geophysical ows and can help
in the hazard prevention for natural disasters.
Acknowledgment
Financial support from EPSRC through grant GR/
R72013 is gratefully acknowledged.
References
[1] Garcia-Navarro P, Priestley A, Hubbard ME. Genuinely
multidimensional upwinding for the 2D shallow water
equations. J Comput Phys 1995;121:7993.
[2] Toro E. Shock capturing methods for free-surface shallow
ows. New York: Wiley; 2001.
[3] Mingham CG, Causon DM. High resolution nite-volume
method for shallow water ows. J Hydr Engrg 1998;124:
60514.
[4] Bathe KJ, Zhang H. Finite element developments for
general uid ows with structural interactions. Int J Num
Meth Engrg 2004;60:21332.
[5] Monaghan JJ. Simulating free surface ows with SPH. J
Comput Phys 1994;110:399406.
[6] Bonet J, Lok T-SL. Variational and momentum preserva-
tion aspects of smooth particle hydrodynamics formula-
tions. Comput Meth Appl Mech Engrg 1999;180:97115.
[7] Monaghan JJ. Particle methods for hydrodynamics. Com-
put Phys Rep 1985;3:7193.
[8] Takeda H, Miyama SM, Sekiya M. Numerical simulation
of viscous ow by smoothed particle hydrodynamics. Prog
Theor Phys 1995;92:939.
[9] Lucy LB. A numerical approach to the testing of the ssion
hypothesis. Astro J 1977;82:1013.
[10] Gingold RA, Monaghan JJ. Smooth particle hydrodynam-
ics: theory and application to non-spherical stars. Mon Not
R Astron Soc 1977;181:375.
[11] Monaghan JJ, Kocharyan A. SPH simulation of multi-
phase ow. Comput Phys Comm 1995;87:22543.
[12] Rodriguez-Paz MX, Bonet J. Mesh-free numerical simula-
tion of debris ows avalanches. Numer Meth Partial Di
Equat 2004;20:14063.
[13] Gutfraind R, Savage SB. Smooth particle hydrodynamics
for simulation of broken-ice elds: MohrCoulomb type
rheology and frictional boundary conditions. J Comput
Phys 1997;134:20315.
[14] Bonet J, Kulasegaram S, Rodriguez-Paz MX, Prot M.
Variational formulation for the smooth particle hydrody-
namics (SPH) simulation of uid and solid problems.
Comput Meth Appl Mech Engrg 2004;193(1214):124556.
[15] Kulasegaram S, Bonet J, Lewis RW, Prot M. A
variational formulation based contact algorithm for SPH
applications. Comput Mech 2004;33(4):31625.
[16] Benz W. Smooth particle hydrodynamics: a review. In:
Buchler JR, editor. The numerical modelling of nonlinear
stellar pulsations. Dordrecht: Kluwer Academic Publish-
ers; 1990. p. 26988.
[17] Mann A. The classical dynamics of particles: Galilean and
Lorentz relativity. New York: Academic Press; 1974.
[18] Rodriguez-Paz MX. Corrected SPH techniques for debris
ows simulations. PhD Thesis, School of Engineering,
University of Wales, Swansea, UK, 2002.
[19] Zienkiewicz OC, Taylor RL. The nite element method:
uid dynamics, vol. 3. London: Butterworth-Heinemann;
2000.
[20] Stoker JJ. Water waves. New York: InterScience; 1965.
[21] Rodriguez-Paz MX, Bonet J. A variational SPH method
for debris ows. In: Wheel M, editor. 11th ACME
Conference. Glasgow: University of Strathclyde; 2003.
[22] Bonet J, Pearaire J. An alternating digital tree (ADT)
algorithm for geometric searching and intersection prob-
lems. Int J Num Meth Engrg 1991:3159.
[23] Swegle JW, Hicks DL, Attaway SW. Smooth particle
hydrodynamics stability analysis. Int J Comp Phys 1995:
11642.
1410 M. Rodriguez-Paz, J. Bonet / Computers and Structures 83 (2005) 13961410

Das könnte Ihnen auch gefallen