Sunteți pe pagina 1din 10

Advances in Water Resources 27 (2004) 259–268

www.elsevier.com/locate/advwatres

Central scheme for two-dimensional dam-break flow simulation


a,* b
G. Gottardi , M. Venutelli
a
Facolta de Ingegneria, Dipartimento di Ingegneria Chimica, Mineraria e delle Tecnologie Ambientali, Universita di Bologna,
Viale del Risorgimento 2, I-40136 Bologna, Italy
b
Dipartimento di Ingegneria Civile, Universita di Pisa, Via Gabba 22, I-56126 Pisa, Italy
Received 21 November 2002; received in revised form 15 August 2003; accepted 30 December 2003

Abstract
A new second-order central scheme (KT), proposed by Kurganov and Tadmor, have been used for the solution of the two-
dimensional dam-break problem. After having assumed as basis of the dam-break phenomenon the set of 2D nonhomogeneous
Saint–Venant equations, the KT scheme in its semi-discrete second-order form has been extended for taking into account the
presence of a source term. Time integration has been performed by using a third-order TVD Runge–Kutta scheme. To demonstrate
the accuracy of this explicit method, the solution obtained by this model for a dam-break problem of one-dimensional flow has been
compared with the corresponding analytical solution, moreover, the solutions obtained for two test cases of two-dimensional flow
have been compared with the experimental results.
 2004 Elsevier Ltd. All rights reserved.

Keywords: Saint–Venant equations; Shallow water flow model; Central difference schemes; Incompressible fluid flow; Shock phenomena; Dam-break
problem

1. Introduction problems was used by Peraire et al. [19], Katopodes and


Wu [14], and Sheu and Fang [20].
In many practical situations of the hydraulic and A large class of shock-capturing schemes has been
environmental engineering it is very important the obtained by using the Godunov method in ‘‘upwind’’
knowledge of the maximum water height resulting from approach (see Hirsch [9]). Different formulations using
a dam-break event caused by a sudden break of a Riemann solvers have been presented, among these, the
holding barrier. This phenomenon is mathematically first-order model using Roe’s numerical flux, and the
described by the set of 2D nonhomogeneous Saint– second-order model, based on Lax–Wendroff numerical
Venant partial differential equations, the solutions of flux have been presented by Jha et al. [10]. Models based
which, excluding some simplified cases, can be obtained on the second-order Lax–Wendroff method have been
only by numerical integration. For the correct repre- presented also by Louaked and Hanich [18] and by
sentation of the dam-break phenomenon the numerical Wang et al. [27]. For the numerical integration of the 2D
schemes used must be able to model the sudden varia- shallow water equations, different Riemann solvers have
tions of the hydraulic parameters without introducing been used also in finite-volume formulations by Alcrudo
spurious oscillations. For this purpose, Katopodes and and Garcia-Navarro [2], Zhao et al. [30], Wang and Liu
Strelkoff [11,12] used the method of characteristics, [28], Tseng and Chu [24] and Brufau and Garcia-
Katopodes [13], and Akanbi and Katopodes [1] used a Navarro [3].
finite-element Petrov–Galerkin formulation, whereas, Recently, Kurganov and Tadmor [15], have presented
Fennema and Chaudhry [5,6] used implicit and explicit a ‘‘central’’ scheme for solving homogeneous convection
finite difference formulations. The Taylor–Galerkin and convection–diffusion equations. Differently from
method, presented by Donea [4] for convective transport the upwind schemes, which compute the reconstructed
values at the mid-cells, the central schemes compute the
staggered cell averages at the interfacing break-points.
*
Corresponding author. Tel.: +39-051-209-3385; fax: +39-051-209-
These central schemes present therefore the advantage
3392. of the simplicity in that no approximate Riemann
E-mail address: guido.gottardi.1@mail.ing.unibo.it (G. Gottardi). solvers are involved in their construction. The principal
0309-1708/$ - see front matter  2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.advwatres.2003.12.006
260 G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268

idea inspiring the KT model is the use of precise infor- In the case of shallow-water the vertical component
mation about the local speed of the wave propagation. of the velocity w and the corresponding shear stress can
In fact, in the KT scheme, an approximate solution is be neglected, in this case Eq. (4c) becomes (see Lai [16])
constructed by using cells the length of which are not pz ¼ qg ð5Þ
constant in time but are computed on the basis of the
local value of the propagation wave celerity. The Integration of (5) gives
evolved solution, after an additional reconstruction, is p ¼ qgðZ  zÞ þ patm ð6Þ
then projected back onto the original grid.
The aim of this work is to present the extension of the where Z ¼ Zðx; yÞ is the water surface elevation (Fig. 1),
semidiscrete KT scheme to the integration of the two- z the vertical coordinate and patm the atmospheric pres-
dimensional Saint–Venant equations in presence of a sure, which is assumed to be constant. From Eq. (6) we
source term, and to verify the efficiency of this formu- obtain
lation for modeling the dam-break phenomenon. The px ¼ qgZx ; py ¼ qgZy ð7a;bÞ
paper is organized as follows: in Section 2 the basic
Another important step for obtaining the set of
equations of the dam-break phenomenon are recalled, in
shallow water equations is the integration along the z
Section 3 the numerical formulation of the model is
coordinate of the continuity Eq. (3) and of momentum
presented and, at last, in Section 4 the results of the
Eqs. (4a) and (4b).
numerical applications of the above described model for
This integration over the depth h, extending from the
one-dimensional idealized problem and for two-dimen-
bottom z ¼ Zb ðx; yÞ to the free surface z ¼ Zðx; y; tÞ,
sional real cases of dam-break phenomena are pre-
where h ¼ Z  Zb , by using the Leibnitz rule and Eq.
sented.
(7a,b), gives the following set of partial differential
equations, which are used for describing the transient
flow in shallow water cases (see Toro [25])
2. Governing equations
ht þ ðhuÞx þ ðhvÞy ¼ 0 ð8aÞ
The mathematical model of an incompressible fluid
ðhuÞt þ ðhu2 Þx þ ðhuvÞy ¼ ghðZx þ Sfx Þ ð8bÞ
flow, in a bounded domain X with boundary C, is based
on the continuity and Navier–Stokes partial differential
equations written in the form ðhvÞt þ ðhuvÞx þ ðhv2 Þy ¼ ghðZy þ Sfy Þ ð8cÞ

ru¼0 ð1Þ The set of conservative Eq. (8), which are known under
the name of 2D Saint–Venant equations, in matrix form
are
ou 1
þ ðu  rÞu ¼  rp þ mr2 u þ f ð2Þ Ut þ Fx þ Gy ¼ S; ð9Þ
ot q
T
where uðx; y; z; tÞ ¼ ðu; v; wÞ is the velocity vector, q
the water density, assumed as a constant, pðx; y; z; tÞ is
the pressure, m the kinematics viscosity, and fðx; y; z; tÞ
is the body force vector. Eqs. (1) and (2), in the equiv-
alent scalar form are
ux þ v y þ w z ¼ 0 ð3Þ

1
ut þ uux þ vuy þ wuz ¼  px þ mr2 u ð4aÞ
q

1
vt þ uvx þ vvy þ wvz ¼  py þ mr2 v ð4bÞ
q

1
wt þ uwx þ vwy þ wwz ¼  pz þ mr2 w  g ð4cÞ
q
here ut ; ux ; uy ; uz indicate, respectively, the partial deriv-
atives ou=ot, ou=ox, ou=oy, ou=oz, the same is true for v
and w and p. The body force vector is f ¼ ð0; 0; gÞ, g
being the acceleration of gravity. Fig. 1. Shallow water notation.
G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268 261

where 3. Kurganov–Tadmor semi-discrete central scheme


2 3 2 3
h uh The hyperbolic set of Saint–Venant partial differential
U ¼ 4 uh 5; F ¼ 4 u2 h þ 12 gh2 5 ð10a;bÞ equations (PDE) (9) can be integrated by using different
vh uvh numerical schemes. In this work the integration of this
set of equation was performed by using the scheme re-
2 3 2 3 cently proposed by Kurganov and Tadmor [15] and used
vh 0
G¼4 uvh 5; S ¼ 4 ghðS0x  Sfx Þ 5 ð10c;dÞ by Gottardi and Venutelli [8], for modeling shock phe-
2 1
v h þ 2 gh 2
ghðS0y  Sfy Þ nomena in open channel flow. With reference to these
papers, it is easy the extension of the formulation given
here S0x  Zbx , S0y  Zby and Sfx and Sfy are, respec- for the integration of the one-dimensional case (1D) of
tively, the bottom slope and the friction slope compo- the Saint–Venant equations to the two-dimensional case
nents in the x and y-directions. The friction slope (2D). The semi-discrete scheme for the 2D Saint–Venant
components according Manning’s formula are Eq. (9), using a uniform spatial grid ðxj ; yk Þ ¼ ðjDx; kDyÞ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffi shown in Fig. 2, reads
un2m u2 þ v2 vn2m u2 þ v2
Sfx ¼ ; Sfy ¼ d
x
Hjþ1=2;k x
 Hj1=2;k
h4=3 h4=3 Uj;k ðtÞ ¼ 
dt Dx
where nm (m1=3 s) is the roughness coefficient. The y y
Hj;kþ1=2  Hj;k1=2
equivalent nonconservative form of Eq. (9) is  þ Ej;k ðtÞ ð14Þ
Dy
Ut þ AUx þ BUy ¼ S ð11Þ
where xj
1=2 :¼ xj
Dx=2 and yk
1=2 :¼ yk
Dy=2, and
where the Jacobian of the fluxes A ¼ oF =oU and the fluxes are
B ¼ oG=oU are
þ 
2 3 2 3 x
F ðUj
1=2;k ðtÞÞ þ F ðUj
1=2;k ðtÞÞ
0 1 0 0 0 1 Hj
1=2;k ðtÞ :¼
2
A ¼ 4 c2  u2 2u 0 5; B ¼ 4 uv v u5
axj
1=2;k ðtÞ

uv v u c2  v2 0 2v  þ
Uj
1=2;k 
ðtÞ  Uj
1=2;k ðtÞ
2
ð12a;bÞ
ð15aÞ
where c ¼ ðghÞ1=2 is the wave speed. The eigenvalues of
A and B are
8 8
>
> a ¼uþc >
> b ¼vþc
< 1 < 1
a
A 2 ¼ u B b ¼v ð13a;bÞ
>
> >
> 2 y k+1
: a3 ¼ u  c : b3 ¼ v  c
 
In addition, to complete the mathematical formulation
of the problem, the following initial conditions y k+1/2

hðx; y; 0Þ ¼ h0 ðx; yÞ; uðx; y; 0Þ ¼ u0 ðx; yÞ

and the appropriate boundary conditions must be yk


specified. The basis for determining the required number (j,k)
of boundary conditions is the local value of the Froude
number or weather the flow is critical or supercritical.
For supercrital flow, no condition at the outflow y k-1/2
boundary and three conditions at the inflow boundary
are necessary. For the subcritical flow, instead, two
boundary conditions, for the inflow and one for the yk - 1
outflow boundary are specified. In these practical cases,
y xj-1 xj x j+1/2
water depth, and velocity or unit discharge at the inflow x j-1/2 x j+1
boundary and depth at the outflow boundary are usualy
specified. Additional boundary conditions, regarding
x
solid walls, no-slip and free-slip boundary conditions
can be imposed (Vreugdenhil [26], and Wesseling [29]). Fig. 2. Two-dimensional grid.
262 G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268

þ  Z xjþ1 Z ykþ1
y
GðUj;k
1=2 ðtÞÞ þ GðUj;k
1=2 ðtÞÞ 1 2 2
Hj;k
1=2 ðtÞ :¼ E2j;k ðtÞ ¼ S2 dx dy
2 DxDy xj1 yk1
ayj;k
1=2 ðtÞ
2 2

þ 
 Uj;k
1=2 ðtÞ  Uj;k
1=2 ðtÞ g
2  hþ 1 S0x  S þ
fxjþ1;k þ h
jþ12;k
S0x  S 
fxjþ1;k
12 jþ2;k
ð15bÞ 2

2

2g   g þ þ

where the intermediate values Uj


1=2;k
ðtÞ and Uj;k
1=2 ðtÞ þ hj;k S0x  Sfxj;k þ h 1 S0x  Sfx 1
3 12 j2;k j ;k
2
are given by the following Taylor expansions
Dx þ h
j1;k

S0x  Sfx ð16bÞ

1
Uj
1=2;k ðtÞ ¼ Uj
1;k ðtÞ ðUx Þj
1;k ðtÞ 2 j ;k
2
2
Z x 1 Z y 1
Dx 1 jþ
2

2

Uj
1=2;k ðtÞ ¼ Uj;k ðtÞ
ðUx Þj;k ðtÞ E3j;k ðtÞ ¼ S3 dx dy
2 DxDy x 1 y 1
j k
2 2


Dy g þ þ  
Uj;k
1=2 ðtÞ ¼ Uj;k
1 ðtÞ ðUy Þj;k
1 ðtÞ  h 1 S0y  Sfy 1 þ hj;kþ1 S0y  Sfy 1
2 12 j;kþ2 j;kþ
2
2 j;kþ
2

2g   g
Dy þ hj;k S0y  Sfyj;k þ hþ S  S þ
Uj;k
1=2 ðtÞ ¼ Uj;k ðtÞ
ðUy Þj;k ðtÞ 3 12 j;k2
1 0y fyj;k1
2
2

and the discrete partial derivatives ðUx Þj;k and ðUy Þj;k þ h S0y  Sfy 1 ð16cÞ
j;k1
were computed by 2 j;k
2

The time integration of the semi-discrete second-or-


ðUx Þj;k
der KT central scheme (14) was performed by means of
Ujþ1;k  Uj;k Ujþ1;k  Uj1;k Uj;k  Uj1;k the third-order TVD Runge–Kutta method, proposed
¼ MM h ; ;h
Dx 2Dx Dx by Shu and Osher [21].

ðUy Þj;k

Uj;kþ1  Uj;k Uj;kþ1  Uj;k1 Uj;k  Uj;k1 4. Numerical applications
¼ MM h ; ;h
Dy 2Dy Dx
To demonstrate the ability of the KT scheme for
where 0 6 h 6 1 and MM denotes the MinMod non- modeling accurately dam-break events, some tests cases
linear limiter function, defined as are presented. At first, the one-dimensional dam-break
in horizontal and frictionless channel, of which is known
8
< minj fxj g if xj > 0; 8j the analytical solution, is presented. Moreover three 2D
MMfx1 ; x2 ; . . .g ¼ maxj fxj g if xj < 0; 8j dam-break test cases are simulated and the results of
: these simulations are compared with experimental data.
0 otherwise:

The local speeds axjþ1;k ðtÞ and ayj;kþ1 ðtÞ are computed by 4.1. One-dimensional dam-break in horizontal and fric-
2 2
n o tionless channel
þ 
axjþ1=2;k ðtÞ ¼ max qðAðUjþ1=2;k ÞÞ; qðAðUjþ1=2;k ÞÞ
In this test we consider an idealized dam-break
n o problem for a rectangular, horizontal and frictionless
ayj;kþ1=2 ðtÞ ¼ max qðBðUj;kþ1=2
þ 
ÞÞ; qðBðUj;kþ1=2 ÞÞ channel, 200 m long. The initial conditions were h1 ¼ 10
m for the liquid depth upstream the dam, which is lo-
where qðAÞ : maxi jai j, in which ai are the eigenvalues cated in the middle of the channel, and h2 ¼ 1 m is the
in Eq. (13a), and qðBÞ   in which b are the
: maxi jbi j, liquid depth downstream the dam. The results of this
eigenvalues in Eq. (13b).  i
simulation are plotted, as function of x, at time t ¼ 6 s in
For the evaluation of the three components Fig. 3: (a) the depth h; (b) the velocity u; (c) the unit
(E1 ; E2 ; E3 ) of the vector E, obtained by integration of discharge q ¼ uh and, (d) the Froude number Fr ¼ u=c.
the three components (S1 ; S2 ; S3 ) of the forcing vector S, Also the corresponding analytical results (see Stoker
we use the Simpson’s quadrature rule and consider as [23]) are plotted in the same figures. For this test we have
constant the bed slopes S0x , S0y in the x- and y-directions used a Courant number Cr ¼ ðu þ cÞDt=Dx ¼ 0:5 and
respectively. In this way we obtain the parameter for computing the numerical derivatives
by the MinMod non-linear limiter function was
E1j;k ðtÞ ¼ 0 for S1 ¼ 0 ð16aÞ h ¼ 1:25. The space domain was discretized by 200 cells.
G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268 263

12 8

h [m] Numerical u [m/s]


Analytical Numerical
7 Analytical
10

8
5

6 4

3
4

2
1

0 0
0 50 100 150 200 0 50 100 150 x [m] 200
x [m]
(a) (b)

30 1.2
2 Fr
q [m /s]
25 1.0

20 0.8

15 0.6

10 0.4

Numerical
Analytical
5 0.2
Numerical
Analytical

0 0.0
0 50 100 150 200 0 50 100 150 200
x [m] x [m]
(c) (d)

Fig. 3. Comparison of the numerical and analytical solutions at t ¼ 6 s: (a) depth h; (b) velocity u; (c) unit discharge q and, (d) Froude number Fr.

These figures show a good agreement between numerical depth in the reservoir and 0.0 m in the channel. The
and analytical solutions. contour line is closed, except the outlet which is consid-
ered open. For this simulation a Manning coefficient
4.2. Two-dimensional dam-break experiment in an L- nm ¼ 0:0095 (m1=3 s) and the bed slopes S0x ¼ S0y ¼ 0
shaped channel were used. The flow domain was discretized by squared
cells 1 · 1 cm. The Courant number Cr ¼ kDt=l, in which
For this test case the flow domain formed by a square k is the maximum among the eigenvalues ai and bi in Eq.
reservoir feeding a L-shaped channel is shown in Fig. 4, (13a,b) and l ¼ Dx ¼ Dy, was taken 0.5,moreover,  we
(Brufau and Garcia–Navarro [3]; Soares Fraz~ ao and used h ¼ 1 as the parameter for computing the numerical
Zech [22]). The initial conditions are zero flow and 0.2 m derivative using the MinMod non-linear limiter function.
264 G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268

The domain is discretized into square elements with


l ¼ 2 cm, moreover we used Cr ¼ 0:5, and h ¼ 1.
In a first series of tests with S0x ¼ S0y ¼ 0, the initial
P6
2.85 depth in the reservoir was taken 0.60 m, and the flood-
plain was considered dry. The contour plot of the depth
reservoir 1.45 at t ¼ 0:5 s is shown in Fig. 8. The observed and simu-
lated depth hydrographs at points: (a) ‘‘O’’ (x ¼ y ¼ 1
2.40 P5
m); (b) ‘‘)5A’’ (x ¼ 0:18, y ¼ 1 m); (c) ‘‘C’’ (x ¼ 0:48,
P1
P2 P3 P4 y ¼ 0:40 m) and, (d) ‘‘9B’’ (x ¼ 1:802, y ¼ 1:45 m are
y 0.45 shown in Fig. 9. In a second series of tests we take
S0x ¼ 7%, S0y ¼ 0, and the initial depth in the reservoir
x 2.40 4.40 of 0.64 m. The contour plot of the depth at t ¼ 0:5 s is
shown in Fig. 10, moreover, in Fig. 11 observed and
Fig. 4. Experimental domain.
simulated depth hydrographs at points (a) ‘‘O’’ and, (b)
‘‘)5A’’ are presented.
We believe that the discrepancies occurring in the first
The contour plot of the depth at t ¼ 1 s is shown in instants of the simulation in Figs. 9(a) and 11(a) are not
Fig. 5. The plots of the simulated and experimental due to numerical approximation but to the type of
hydrographs, that were obtained by Prof. Y. Zech, S. mathematical model, that is, the shallow water formu-
Soares and their coworkers at the Laboratory of the lation used for this simulation.
Civil Engineering of the Universite Catholique de Lou-
vain (Belgium) and were handed over by Brufau (Brufau
private communication, 2001) are shown in Fig. 6, 4.4. Two-dimensional circular dam-break experiment
respectively, at points: (a) ‘‘P1’’ (1.20, 1.20 m); (b) ‘‘P2’’
(2.75, 0.70 m); (c) ‘‘P3’’ (4.25, 0.70 m); (d) ‘‘P4’’ (5.75, In this test case, we consider a circular dam in ide-
0.70 m); (e) ‘‘P5’’ (6.55, 1.50 m) and, (f) ‘‘P6’’ (6.55, 3.00 alized and horizontal bottom and study the time evo-
m). lution of the shock waves associated with the breaking
of the dam.
The flow domain was discretized by a grid of 50 · 50
4.3. Two-dimensional dam-break experiment of Fraccar- cells with each cell 1 · 1 m. The initial conditions are
ollo and Toro u ¼ v ¼ 0 throughout the domain, whereas the water
depth inside and outside the dam is respectively, h0 ¼ 10
The experimental flume, shown in Fig. 7, is 2 m wide m and h0 ¼ 1 m (Fig. 12). The plot of the free surface
and 3 m long, 1 m of which is occupied by the reservoir elevation at t ¼ 0:69 s is shown in Fig. 13. The contours
(see Fraccarollo and Toro [7]). The width of the gate, of the water surface elevations at t ¼ 0:69 s are shown in
symmetrically centred, is 0.40 m. The flood-plain Fig. 14. From these figures the perfect symmetry of the
boundaries are all open. In the simulation model flow, as in [2,24], is evident. In this test we have assumed
velocities normal to closed boundary are taken equal to Cr ¼ 0:5 and h ¼ 1. Water depth and Froude number,
zero, whereas, in the outflow open boundary longitudi- along the line O–C in Fig. 12, are shown, respectively, in
nal gradients of u, v and h were assumed to zero. Fig. 15(a) and (b). These results are in good agreement
with those recently obtained by Lin et al. [17] by using
finite-volume TVD shemes.

5. Conclusions

The application of the KT method, recently proposed


by Kurganov and Tadmor, for the integration of the De
Saint–Venant equations, in 2D and in presence of a
0.20 source term due to bottom and friction slopes, has been
presented. The integration in time of the semidiscrete
0.02
0.18 second-order KT formulation has been obtained by a
0 TVD third order Runge–Kutta scheme. From the results
0.0
of the numerical dam-break experiments we can deduce
Fig. 5. Contour plot of the depth at t ¼ 1 s. the following conclusions:
G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268 265

0.20 0.20
h [m] Numerical h [m]
Experimental Numerical
Experimental
0.15 0.15

0.10 0.10

0.05 0.05

0.00 0.00
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
t [s] t [s]
(a) (b)

0.20 0.20
h [m] Numerical h [m] Numerical
Experimental Experimental
0.15 0.15

0.10 0.10

0.05 0.05

0.00 0.00
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
t [s] t [s]
(c) (d)

0.20 0.20
h [m] Numerical h [m] Numerical
Experimental Experimental
0.15 0.15

0.10 0.10

0.05 0.05

0.00 0.00
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
t [s] t [s]
(e) (f)

Fig. 6. Hydrographs at points: (a) P1; (b) P2; (c) P3; (d) P4; (e) P5 and, (f) P6.

0.55
2.0
reservoir flood-plain
[m]
9B

0.50 0.50
-5A 0.05
1.0 0 0.4 m

C
y

0.0 0.00
0.0 x 1 .0 2.0 3.0
[m]
Fig. 8. Contour plot of the depth at t ¼ 0:5 s, for the case
Fig. 7. Experimental reservoir and flood-plain. S0x ¼ S0y ¼ 0.
266 G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268

0.8 0.8
h [m] Numerical h [m] Numerical
Experimental Experimental
0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0 2 4 6 8 10 0 2 4 6 8 10
t [s] t [s]
(a) (b)

0.8 0.8
h [m] Numerical h [m] Numerical
Experimental Experimental
0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0 2 4 6 8 10 0 2 4 6 8 t [s] 10
t [s]
(c) (d)

Fig. 9. Hydrographs for the case S0x ¼ S0y ¼ 0 at points: (a) O; (b) )5A; (c) C; (d) 9B.

0
0.8
h [m] Numerical
Experimental
0.6
0.60

0.4
0.55 0.05
0.50
0.2
0.55

0.60 0.0
0 2 4 6 8 t [s] 10
(a)
0.00
0.8
Fig. 10. Contour plot of the depth at t ¼ 0:5 s for the case S0x ¼ 0:07 h [m] Numerical
and S0y ¼ 0. Experimental
0.6

0.4
1. The presented 1D dam-break test for the frictionless
channel shows a very good agreement among the 0.2
numerical and analytical solutions.
2. For that it concerns the 2D dam-break test cases pre- 0.0
0 2 4 6 8 10
sented, the lack of 2D analytical solutions, to be used t [s]
for comparison, do not allow a certain evaluation of (b)
the goodness of the 2D KT method presented. How- Fig. 11. Hydrographs for the case S0x ¼ 0:07 and S0y ¼ 0, at points: (a)
ever, the validity of the 1D KT method, and the good O and, (b) )5A.
G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268 267

50 0

h0 = 1 m
y [m] 1.01

2.00

4.00
6.00
8.00
C

y [m]
O 9.58
25 25
h0 = 10 m

0 50
0 25 50 0 25 50
x [m]
x [m]
Fig. 12. Flow domain with initial water depth. Fig. 14. Contour plot of the water depth at t ¼ 0:69 s.

12
h [m]
y
10 KT
40 50 LLFS
10 20 30
0
10 8

7.5 6

h 5
4
2.5
2
0
0 0
10
20 25 30 35 40 45 50
30
40 x [m]
50 (a)
x

Fig. 13. Water surface elevation at t ¼ 0:69 s. 2.0

Fr KT
LLFS
agreement among 2D experimental and simulated re- 1.5
sults led to the conclusion of the validity of the 2D
KT computational model.
1.0

Acknowledgements
0.5

The authors thank Prof. P. Brufau of the Centro


Politecnico Superior, Universidad de Zaragoza, Spain, 0.0
for the kind concession of the experimental data ob- 25 30 35 40 45 50
tained by Prof. Y. Zech, S.Soares and their coworkers at x [m]
(b)
the Department of Civil Engineering of the Universite
Catholique de Louvain (Belgium), regarding the dam- Fig. 15. Water depth (a), and Froude number (b), along line O–C, at
break test in a L-shaped channel. t ¼ 0:69 s.
268 G. Gottardi, M. Venutelli / Advances in Water Resources 27 (2004) 259–268

References [16] Lai C. Numerical modeling of unsteady open-channel flow. Adv


Hydros 1986;14:161–333.
[1] Akanbi A, Katopodes N. Model for flood propagation on initially [17] Lin GF, Lai JS, Guo WD. Finite volume component-wise TVD
dry land. J Hydraulic Eng ASCE 1988;114(7):689–706. schemes for 2D shallow water equations. Adv Water Resour
[2] Alcrudo F, Garcia-Navarro P. A high-resolution Godunov-type 2003;26:861–73.
scheme in finite volumes for the 2D shallow-water equations. Int J [18] Louaked M, Hanich L. TVD scheme for the shallow water
Num Meth Fluids 1993;16:489–505. equations. J Hydraulic Res 1998;36:363–78.
[3] Brufau P, Garcia-Navarro P. Two-dimensional dam break flow [19] Peraire J, Zienkiewicz OC, Morgan K. Shallow water problems: a
simulation. Int J Num Meth Fluids 2000;33:35–57. general explicit formulation. Int J Num Meth Engng 1986;22:547–
[4] Donea J. A Taylor–Galerkin method for convective transport 74.
problems. Int J Num Meth Engng 1984;20:101–19. [20] Sheu TWH, Fang CC. High resolution finite-element analysis of
[5] Fennema RJ, Chaudhry MH. Implicit methods for two-dimen- shallow water equations in two dimensions. Comput Meth Appl
sional unsteady free-surface flows. J Hydraulic Res 1989;27:321– Mech Engng 2001;190:2581–601.
32. [21] Shu CW, Osher S. Efficient implementation of essentially non-
[6] Fennema RJ, Chaudhry MH. Explicit methods for 2-D transient oscillatory shock-capturing schemes. J Comp Phys 1988;77:439–
free-surface flows. J Hydraulic Eng ASCE 1990;116(8):1013– 71.
34. [22] Soares Fraz~ao S, Zech Y. Dam Break in channels with 90 bend. J
[7] Fraccarollo L, Toro EF. Experimental and numerical assessment Hydraulic Eng ASCE 2002;128(11):956–68.
of the shallow water model for two-dimensional dam-break type [23] Stoker JJ. Water waves. New York: Interscience Publishers; 1957.
problems. J Hydraulic Res 1995;33:843–64. [24] Tseng MH, Chu CR. Two-dimensional shallow water flows
[8] Gottardi G, Venutelli M. Central schemes for open-chaannel flow. simulation using TVD-MacCormak scheme. J Hydraulic Res
Int J Num Meth Fluids 2003;41:841–61. 2000;38:123–31.
[9] Hirsch C. Numerical computation of internal and external flows. [25] Toro EF. Shock-capturing methods for free-surface shallow flows.
Chichester: John Wiley & Sons; 1995. New York: Wiley; 2001.
[10] Jha AK, Akiyama J, Ura M. Flux-difference splitting schemes for [26] Vreugdenhil CB. Numerical methods for shallow-water flow.
2D flood flows. J Hydraulic Eng ASCE 2000;126(1):33–42. Dordrecht: Kluver; 1994.
[11] Katopodes N, Strelkoff T. Computing two-dimensional dam- [27] Wang JS, Ni HG, He YS. Finite-difference TVD scheme for
break flood waves. J Hydraulic Eng ASCE 1978;104(9):1269–88. computation of dam-break problems. J Hydraulic Eng ASCE
[12] Katopodes N, Strelkoff T. Two-dimensional shallow water-wave 2000;126(4):253–62.
models. J Mechanic Eng ASCE 1979;105(2):317–34. [28] Wang JW, Liu RX. A comparative study of finite volume methods
[13] Katopodes N. Two-dimensional surges and shocks in open on unstructured meshes for simulation of 2D shallow water wave
channels. J Hydraulic Eng ASCE 1984;110(6):794–812. problems. Math Comput Simul 2000;53:171–84.
[14] Katopodes N, Wu CT. Explicit computation of discontinuous [29] Wesseling P. Principles of computational fluid dynamics. Berlin:
channel flow. J Hydraulic Eng ASCE 1986;112(6):456–75. Springer; 2001.
[15] Kurganov A, Tadmor E. New high-resolution central schemes for [30] Zhao DH, Shen HW, Lai JS, Tabios GQ. Approximate Riemann
nonlinear conservation laws and convection–diffusion equations. J solvers in FVM for 2D hydraulic shock wave modeling. J
Comp Phys 2000;160:241–82. Hydraulic Eng ASCE 1996;122(2):692–702.

S-ar putea să vă placă și