Sei sulla pagina 1di 35

WATER RESOURCES RESEARCH, VOL. ???, XXXX, DOI:10.

1029/,

An Exact Solution for Ideal Dam-Break Floods on


Steep Slopes
C. Ancey, M. Rentschler,

School of Architecture, Civil and Environmental Engineering, Ecole

Polytechnique Federale de Lausanne, 1015 Lausanne, Switzerland


R.M. Iverson, R.P. Denlinger

U.S. Geological Survey, Cascades Volcano Observatory, Vancouver,

Washington, USA

C. Ancey, School of Architecture, Civil and Environmental Engineering, Ecole


Polytechnique
Federale de Lausanne, 1015 Lausanne, Switzerland (christophe.ancey@epfl.ch)

D R A F T

September 26, 2007, 3:54pm

D R A F T

X-2

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

Abstract.

The shallow-water equations are used to model the flow re-

sulting from the sudden release of a finite volume of frictionless, incompress-

ible fluid down a uniform slope of arbitrary inclination. The hodograph trans-

10

formation and Riemanns method make it possible to transform the govern-

11

ing equations into a linear system and then deduce an exact analytical so-

12

lution expressed in terms of readily evaluated integrals. Although the solu-

13

tion treats an idealized case never strictly realized in nature, it is uniquely

14

well-suited for testing the robustness and accuracy of numerical models used

15

to model shallow-water flows on steep slopes.

D R A F T

September 26, 2007, 3:54pm

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

X-3

1. Introduction
16

Dam-break floods on steep slopes occur in diverse settings. They may result from

17

failure of either natural or manmade dams, and they have been responsible for the loss

18

of thousands of lives [Costa, 1988]. Recent disasters resulting from dam-break floods on

19

steep slopes include those at Fonte Santa mines, Portugal, in November 2006 and Taum

20

Sauk, Missouri, USA, in December 2005.

21

Numerical solutions of the shallow-water equations are generally used to predict the

22

behavior of dam-break floods, but exact analytical solutions suitable for testing these

23

numerical solutions have been available only for floods with infinite volumes, horizontal

24

beds, or both [e.g., Zoppou and Roberts, 2003]. Computational models used to simulate

25

dam-break floods commonly produce numerical instabilities and/or significant errors close

26

to the moving front when steep slopes and/or irregular terrain are present in the flood

27

path. In part these problems reflect the complex interaction of phenomena not included

28

in model formulation (e.g., intense sediment transport under time-dependent flow con-

29

ditions), but in part they also reflect shortcomings in the numerical solution algorithms

30

themselves. Therefore, it is important to obtain exact analytical solutions of the shallow-

31

water equations that can be used to test the robustness of numerical models when they

32

are applied to floods of finite volume on steep slopes. This paper presents a new solution

33

for this purpose.

34

For the dam-break problem on a horizontal bed, many exact and approximate analyt-

35

ical solutions already exist. For example, Ritter [1892] addressed the case of an infinite

36

volume of fluid suddenly released on a frictionless plane. An exact solution for a dam-

D R A F T

September 26, 2007, 3:54pm

D R A F T

X-4

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

37

break flood of finite volume on a frictionless bed was not presented until Hogg [2006]

38

analyzed the finite-volume lock-exchange problem. The more realistic case involving a

39

rough bed (represented by a Chezy-like friction force) has been addressed by a number of

40

authors, including Whitham [1954], Dressler [1952], and Hogg and Pritchard [2004], but

41

only asymptotic solutions have been developed to date. Taking into account a nonuniform

42

velocity distribution in the vertical direction leads to mathematical difficulties, but exact

43

self-similar solutions can still be obtained for floods with variable inflow (i.e., the released

44

volume is a function of time) [Ancey et al., 2006, 2007].

45

For sloping beds, most dam-break solutions developed to date employ approximations

46

of the shallow-water equations, in which inertia or pressure-gradient terms have been

47

neglected. Such assumptions typically lead to a kinematic wave approximation, which en-

48

ables substantial simplification because the mass and momentum balances making up the

49

shallow-water equations are transformed into a single nonlinear diffusion equation [Hunt,

50

1983; Daly and Porporato, 2004a, b; Chanson, 2006]. Exact solutions of the shallow-water

51

equations for steep slopes have been obtained for infinite-volume dam-break floods [Shen

52

and Meyer , 1963; Mangeney et al., 2000; Karelsky et al., 2000; Peregrine and Williams,

53

2001], and the case of a finite-volume flood has been investigated by Dressler [1958] and

54

later by Fernandez-Feria [2006], who provided a partial solution by computing the po-

55

sition and velocity of the surge front and rear.

56

two similarity solutions known as the parabolic cap and M-wave, but these differ from the

57

long-time asymptotic solution of the problem investigated here.

Savage and Hutter [1989] constructed

58

In this paper we present a new analytical solution of the shallow-water equations for a

59

situation in which a finite volume of an ideal (frictionless) fluid is instantaneously released

D R A F T

September 26, 2007, 3:54pm

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

X-5

60

from behind a dam on a steep slope. Although frictionless flows never occur in real fluids,

61

the frictionless case constitutes an unambiguous end member as well as a clear target case

62

for testing numerical models [Zoppou and Roberts, 2003]. Our solution strategy is mostly

63

identical to that used by Hogg [2006] for the lock-exchange problem, with some additional

64

complications that we shall detail later. We begin our analysis by using the characteristics

65

of the shallow-water equations to infer the positions of the flow front and tail at all

66

times. We then employ the hodograph transformation, which converts the nonlinear

67

shallow-water equations into a linear system by exchanging the roles of the dependent

68

and independent variables. An integral form of the exact solution of the linear equations

69

is then obtained using Riemanns method. This method, seldom used in open-channel

70

hydraulics, is well established in some other fields where hyperbolic equations similar to

71

the shallow-water equations arise. Typical examples include gas dynamics [Courant and

72

Friedrich, 1948], collapse of a granular column [Kerswell , 2005], and tsunami or swash

73

run-up on a shore [Carrier and Greenspan, 1958].

2. Governing equations
2.1. Flow-depth averaged equations
74

The nonlinear, one-dimensional shallow-water (Saint-Venant) equations provide a suit-

75

able approximation for modeling water surges over a wide, uniformly sloping bed inclined

76

at an angle with respect to the horizontal (Figure 1). If the effects of friction are

77

neglected (see Appendix A), these equations may be written as


(h
u) = 0,
h+
x
t


u + u u + g cos h
= g sin ,

x
x
t

D R A F T

September 26, 2007, 3:54pm

(1)
(2)

D R A F T

X-6

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

78

where x is the downstream coordinate, t is time, g is the magnitude of gravitational

79

x, t) is the flow depth


acceleration, u(
x, t) is the depth-averaged flow velocity, and h(

80

measured perpendicular to the bed. Note that we use the shallow-water equations in a

81

non-conservative form, which is permitted since the solution to the initial-boundary-value

82

problem investigated here is smooth. Originally, the Saint-Venant equations were derived

83

to model flood propagation on shallow slopes and smooth topography [Saint Venant,

84

1871], but modern formulations have demonstrated that the equations can be recast to

85

apply rigorously to steep slopes and irregular topography [Dressler , 1978; Bouchut et al.,

86

2003; Keller , 2003].


Equations (12) can be normalized using the following scaled variables
x
x=
, t=
H0

g cos
h
u
t, h =
, and u =
,
H0
H0
gH0 cos

87

where H0 is the initial fluid depth at the dam wall. Substitution of the scaled variables

88

into (1) and (2) yields the following dimensionless equations


h
h
u
+u
+h
= 0,
t
x
x
u
u h
+u
+
= tan ,
t
x x

89

(3)
(4)

which can be recast in the matrix form

U+A
U = B,
t
x

90

91

92

with

"

h
,A=
u

"

u h
, and B =
1 u

"

0
U=
.
tan

The matrix A has two real eigenvalues given by = u h, indicating that the shallow
water equations are fully hyperbolic and that h can be identified as the dimensionless

D R A F T

September 26, 2007, 3:54pm

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

93

wave celerity, c =

94

their characteristics as [Stoker , 1957; Whitham, 1974; Chanson, 2004]

h. The hyperbolic system of equations can be expressed in terms of

d
dx
= tan along the -characteristic curve:
= u + c,
dt
dt
95

(5)

where = u + 2c is the associated Riemann variable; and


d
dx
= tan along the -characteristic curve:
= u c,
dt
dt

96

X-7

(6)

where = u 2c the other Riemann variable.


2.2. Initial and boundary conditions for the dam-break problem

97

We consider a situation in which a dam perpendicular to the slope initially retains a

98

reservoir behind it, as shown in Figure 1. The reservoir geometry is defined in cross

99

section by the triangle OAB, where OA denotes the dam wall. The initial water depth is

100

h = h0 (x) = 1 x/xb , where xb = 1/ tan represents the abscissa of point B in Figure 1.

101

At time t = 0, the dam collapses instantaneously and unleashes a flood of finite volume

102

down the slope. An important difference between our formulation and that of Fernandez-

103

Feria [2006] lies in the initial configuration of the flow, because Fernandez-Feria [2006]

104

investigated the case of a vertical dam. Although a vertical dam is more similar to some

105

real-world scenarios, it leads to significant mathematical difficulties when the method of

106

characteristics is employed owing to singular behavior of the front and rear (both u and

107

h being zero there).

108

Following the dam break, part of the water immediately moves downstream in the form

109

of a forward wave, while a wave propagating upstream separates moving fluid from static

110

fluid upslope. The downstream and upstream waves constitute moving boundaries issuing

111

from the origin point in the x t plane (Figure 2). One boundary corresponds to the flow

D R A F T

September 26, 2007, 3:54pm

D R A F T

X-8

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

112

front, where h = 0 and u = uf (uf being the front velocity, unknown for the present). The

113

other boundary constitutes the locus of the upstream propagating wave, which travels to

114

point B in Figure 1. Along this wave, we have h = h(t) (which is also unknown at present)

115

and u = 0.

116

Mathematically, the two moving boundaries are described by characteristic curves in

117

the x t plane, which can be computed using (5) or (6), with h = 0 (forward front)

118

and u = 0 (backward wave). For the forward wave, equation (5) reduces to du/dt =

119

tan . The initial condition applicable with this equation is u = 2 at t = 0 because the

120

dam collapse theoretically causes instantaneous acceleration at t = 0 such that the front

121

velocity immediately becomes u = 2, independently of slope. Although this instantaneous

122

acceleration appears unrealistic physically, it is a logical consequence of the shallow-water

123

approximation, and it can be demonstrated mathematically by noting that the initial

124

value of the Riemann variable is 2; at early times after the dam collapse, since the flow-

125

front depth drops to zero, this value implies that u = 2 at the flow front. Use of this value

126

as the initial condition in du/dt = tan yields the front velocity solution u = t tan + 2.

127

Moreover, because u = dx/dt, we deduce that x = 21 t2 tan + 2t is the locus of the front

128

position in the x t plane.

129

To obtain the speed of the wave that propagates upstream from the dam into still water,

130

we infer from (6) that d(2c)/dt = tan along the characteristic curve. Integration of

131

this equation gives c = tan2 t + 1 since at t = 0, we have c = 1. Substitution of this

132

result into the equation defining the characteristic, dx/dt = c yields x =

133

the equation governing propagation of the backward wave in the x t plane. According

D R A F T

September 26, 2007, 3:54pm

tan 2
t
4

t as

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

X-9

134

to this equation, point B in Figures 1 and 2 is reached by the backward wave at time

135

tb = 2cotan.

136

Once point B is reached, a new wave issues from point B and defines the speed of the

137

moving tail of the volume of fluid as it descends the slope. Propagation of this wave

138

follows the trajectory BC in Figure 2. At the tail margin, the condition h = 0 (c = 0)

139

applies, just as at the front of the forward wave. At point B, the initial conditions for

140

the characteristic equation are x = xb = cotan, t = tb = 2 cotan, h = 0 and u = 0.

141

Substituting c = 0 in (5) and integrating the resulting equation du/dt = tan yields the

142

wave velocity u = tan (t tb ) = t tan 2. Integrating this equation once again yields

143

the equation describing the position of the moving tail in the x t plane:

1 2
t2
x = tan
t ttb + b
2
2

+ xb =

t2
tan 2t + cotan.
2

144

Tables 1 and 2 summarize all the equations defining the boundaries of the moving fluid,

145

and Figure 2 illustrates the position of the boundaries in the x t plane.

146

Some key physical implications of the boundary equations listed in Tables 1 and 2 de-

147

serve special mention. First, once motion of the head and tail begins from their respective

148

initial conditions, each boundary propagates downslope with an acceleration identical to

149

that of a frictionless point mass moving along the slope. This finding implies that the

150

boundary speeds are uninfluenced by the presence of adjacent fluid after motion com-

151

mences. Second, the speed of the advancing flow front always exceeds that of the ad-

152

vancing tail by 4 for t > tb . The difference in speeds is inherited from the difference in

153

initial conditions affecting the head and tail, and it implies that the traveling wave of fluid

154

continuously elongates at a constant rate. This constant elongation would not occur, of

155

course, in a flow with frictional dissipation.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 10

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

3. Homogenization and hodograph transformation


156

In order to make the governing equations homogeneous and simplify calculations, we

157

use a change in variables so that the effects of gravitational acceleration do not appear

158

explicitly:
tan 2
= h,
= x
t , t = t, v = u t tan , and h
2

159

(7)

Use of these substitutions in (3) and (4) yields


h
h
v
+v
+h
= 0,
t

v
v h
+v
+
= 0,
t

(8)
(9)

160

where the tilde has been removed to simplify notation. The characteristic form of these

161

equations is now
dr
d
= 0 along the r-characteristic curve:
= v + c,
dt
dt

162

where r = v + 2c is a Riemann invariant, and


ds
d
= 0 along the s-characteristic curve:
= v c,
dt
dt

163

(10)

(11)

where s = u 2c is the other Riemann invariant.

164

The next step is linearization in order to use analytical methods available for linear

165

partial differential equations [Garabedian, 1964]. Transformation of the governing equa-

166

tions into quasi-linear equations is made possible by using hodograph variables. That is,

167

instead of seeking solutions in the form h(, t) and v(, t), we switch the dependent and

168

independent variables and seek solutions in the form (v, h) and t(v, h) or, more precisely,

169

(r, s) and t(r, s) since we have

1
1
v = (r + s) and h = (r s).
2
4
D R A F T

September 26, 2007, 3:54pm

(12)
D R A F T

X - 11

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED


170

Denoting the Jacobian of the transformation by J = h tv v th , we obtain


h =

tv
th
v
h
, v = ht = , and vt = .
J
J
J
J

171

The transformation is reversible provided J 6= 0 and 1/J 6= 0. This condition is satisfied

172

here except at the flow boundaries, but since the solution is known there (as summarized in

173

Table 2), this restriction presents no difficulty. With the new variables, the homogeneous

174

governing equations (8) and (9) reduce to

175

176

t
t
+v
h
= 0,
v
v
h

t
t
+
v
= 0.
h v
h

(13)
(14)

Equations (13) and (14) can be solved using the method of characteristics. The equation
of an r-characteristic in the plane r s is given by

3r + s t
=
,
s
4 s

(15)

177

which was deduced from Eq. (10) using d = s ds and dt = ts ds since r is constant.

178

Similarly, we obtain for the s-characteristic equation

3s + r t
=
.
r
4 r

(16)

179

We next derive a single equation governing t. Differentiating Eq. (15) with respect to r

180

and Eq. (16) with respect to s, then finding the difference of the two resulting equations,

181

we obtain the equation for t:


3
2t

L[t] = 0 where L[t] =


rs 2(r s)

t
t

.
r s

(17)

182

A similar equation can be obtained for , but its form is more complicated and it is more

183

fruitful to compute t by solving Eq. (17) and then using one of the characteristic equations

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 12

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

184

(15) or (16) to find . Equation (17) is a linear hyperbolic partial differential equation of

185

second order, which arises in a number of contexts in gas dynamics and hydrodynamics

186

and for which solutions are known in terms of Riemann functions [Courant and Friedrich,

187

1948; Garabedian, 1964; Kevorkian, 2000]. The boundary conditions for Eq. (17) are

188

specified along curves OA, OB, and BC (see Table 1).

4. Riemann formulation
189

Next we exploit the linearity of Eq. (17) and use an integral representation to relate t

190

to its auxiliary conditions. If we integrate Eq. (17) over a finite domain D whose oriented

191

contour is denoted by , we obtain area integrals that by themselves yield little insight.

192

However, if we transform these area integrals into boundary integrals using Greens theo-

193

rem, then part of the problem is solved. In this context, Riemanns formulation involves

194

introducing an adjoint differential operator N ( ), which enables us to write [Garabedian,

195

1964; Zauderer , 1983]


L[t] tN [ ] = U =

196

U
V
+
,
r
s

where U = (U, V ) is a vector field. In this way, we obtain


Z

( L[t] tN [ ])drds =

U nd,

(18)

197

where n is an outward normal vector along and d is a curvilinear abscissa such that

198

nd = (ds, dr). For this decomposition to hold, we must define N , U , and V as follows

N [ ] =
rs 2(r s) r
s
t
t
3 1
t +

,
U =
2rs
2 s 2 s
3 1
t
t
V =
t +

.
2rs
2 r 2 r

D R A F T

3
.
(r s)2

September 26, 2007, 3:54pm

(19)
(20)
(21)

D R A F T

X - 13

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED


199

We now consider a geometric domain D in the form a quadrilateral MPOQ, as depicted

200

in Fig. 3. The value of t is known along PO (point O corresponds to point O in the x t

201

plane) and OQ (see Table 1). Since we are free to choose the function , we pose
N [ ] = 0,

202

(22)

with the boundary conditions


(a, b) = 1,

3
=
on r = a, and
=
on s = b,
s
2(r s)
r
2(r s)

(23)

203

These equations remove the dependency on v in the boundary integrals along PM and

204

QM. The solution of (22) satisfying these boundary conditions may be written as the

205

Riemann function R(r, s ; a, b):


"

3 3
(r s)3
(r a)(s b)
.
(r, s) = R(r, s ; a, b) =
F , , 1,
3/2
3/2
(r b) (s a)
2 2
(r b)(s b)

(24)

206

where F is the hypergeometric function [Abramowitz and Stegun, 1964, p. 556]. A deriva-

207

tion of (24) is provided in Appendix B.

208

Identifying the function as in (24) and making use of (22), (18) becomes

Und = 0.

209

The oriented contour line can be broken down into segments QM and MP, where the

210

boundary conditions (23) hold, and the segments PO and OQ (Figure 3), leading to
Z

U nd =

Z M
Q

V dr +

Z P
M

U ds

Z O
P

V dr +

Z Q
O

U ds = 0.

(25)

211

After integrating the boundary integrals in (25) by parts and making use of (23), we

212

rearrange the contribution along each segment of as follows

Z M

1
1
1
3
+
t

dr = t(Q) (Q) + t(a, b),


V dr = [t ]M
Q
2
2 r b r
2
2
Q
Q
!
Z P
Z P
1
3

1
1
U ds = [t ]PM +
t

ds = t(P ) (P ) t(a, b),


2
2 a s s
2
2
M
M

Z M

D R A F T

September 26, 2007, 3:54pm

(26)
(27)

D R A F T

X - 14

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

Z O

Z O
1
3 t
t
V dr = [tR]O
+
R(r, s ; a, b)
+
dr,
P
2
2 r + 2 r
P
P

!
Z Q
Z Q
1
3 t
t
Q
U ds = [tR]O +
R(r, s ; a, b)
+
ds.
2
2 2 s s
O
O

(28)
(29)

213

In so doing, we obtain from the right-hand sides of (26)(29) an integral representation

214

of t that holds for any point M (a, b) inside the triangle OFB in the r s plane
Z Q
1
1
(U ds V dr).
t(a, b) = t(P )R(P ; M ) + t(Q)R(Q ; M ) +
2
2
P

215

216

Since on the boundaries PO and OQ we have tr = cotan/2 and ts = 0, respectively,


we can reduce the equation for t to
t(a, b) = cotan

217

Z a
2

R(r, 2 ; a, b)

2 5r
dr.
4(r + 2)

The variable is then computed by integrating an s-characteristic, i.e., Eq. (16)


1
1Z 2 0
(r|s = cst) = (3s + r)t(r, s) +
t(r , s)dr0 ,
4
4 r

218

(30)

(31)

where we have taken into account the boundary condition = 0 at t = 0.

219

Although equations (30) and (31) are not fully explicit expressions, these exact integral

220

solutions can be evaluated numerically without any difficulty by using computing software

221

such as Mathematica. The Mathematica notebook used to plot the figures in this paper is

222

available online from our website (http://lhe.epfl.ch). The solutions can also be expressed

223

in terms of Legendre functions and computed using tabulated values. Note that when

224

0, time t tends toward infinity, which means that with this solution, we cannot recover

225

the solution calculated by Hogg [2006] for a horizontal plane. This restriction results from

226

the differing upstream boundary condition in the two problems. For a horizontal bottom,

227

part of the fluid remains in the reservoir and the velocity at point B is zero, whereas for a

D R A F T

September 26, 2007, 3:54pm

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

X - 15

228

sloping bed, once the backward wave has reached the upstream end of the reservoir, the

229

tail of the flood wave starts moving and its velocity is nonzero (see Table 1).

230

Equations (30) and (31) form an implicit solution to Eqs. (89) that can be quite easily

231

inverted to provide h(, t) and v(, t). Figure 4 shows the s- and r-characteristics obtained

232

when the bed slopes at the angle = /4. Figure 5 shows the flow-depth and velocity

233

profiles at different times after the dam collapse for = /4. The graphs of Figure 5

234

depict the flow depth and velocity profiles in a frame moving at velocity t tan . Note

235

that the velocity variations are nearly linear and the flow depth profile is increasingly

236

symmetric as elapsed time increases. These features are reminiscent of the parabolic-cap

237

similarity solution of Savage and Hutter [1989]. A shown in Appendix C, however, the

238

parabolic cap solution differs from the long-time asymptotic solution of the shallow-water

239

equations we present here.

240

Expressing our solution in terms of the original dimensionless variables x and t is

241

straightforward. The value of x is given by x = + 12 tan t2 , while t remains unchanged.

242

Figure 6 uses these variables to depict the flow-depth and velocity profiles at different

243

times after the dam collapse, and figure 7 shows details of the evolution of flow depth

244

at early times. Combining the velocity and flow depth profiles at early times makes it

245

possible to evaluate the discharge at the dam site and thereby to obtain a hydrograph

246

that can be used to provide initial conditions in numerical models that route floods using

247

the shallow-water equations. Finally, note that the shape of the characteristic curves in

248

the x t plane is significantly altered due to fluid acceleration. Figure 8 shows the - and

249

-characteristics in the x t plane for = /4.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 16

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

251

can be represented parametrically by using the


The physical variables x, t, u, and h

dimensionless auxiliary variables r = v + 2 h and s = v 2 h (i.e., the Riemann

252

invariants)

250

Z r
H0
H0
2 5
R(, 2 ; r, s)
t(r, s) =
t=
cotan
d,
g cos
g cos
4( + 2)
2

Z 2
q
q
1
2

x(r, s) = H0 x =
(3s + r) gH0 cos t + gH0 cos
t(, s)d + 2 sin g t ,
4
r
q

u(r, s) =

(32)
(33)

gH0 cos u = v gH0 cos + g tsin ,

(34)

s) = H0 h,
h(r,

(35)

253

for r > s > 2 and 2 < r < 2 and where R is the Riemann function given by Eq. (24).

254

For s = 2 and 2 < r < 2, which apply to the backward wave for 0 < t < tb , we have
s

t(r, 2) =

H0
r
1
cotan,
g cos
2

(36)

2) = H0 1 t g cos tan .
h(r,
H0

(37)

255

The case r = s (with s > 2) corresponds to t , while r = 2 (with s > 2)

256

corresponds to the initial condition before the dam breaks. The particular value r = s = 2

257

gives the position and velocity of the flow front, while r = s = 2 gives the position and

258

velocity of the flow tail after the fluid has detached from point B (i.e., for t > tb ):
q

2) = 0 and u(2, 2) = g tsin + 2 gH0 cos ,


h(2,

(38)

h(2,
2) = 0 and u(2, 2) = g tsin 2 gH0 cos .

(39)

5. Conclusion
259

By employing the one-dimensional shallow-water equations, an accelerated reference

260

frame, hodograph transformation, and Riemanns method, we have derived a new exact

261

solution describing the behavior of a dam-break flood of finite volume traveling down
D R A F T

September 26, 2007, 3:54pm

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

X - 17

262

a steep, planar slope. Although the solution assumes that the fluid is frictionless, it

263

nonetheless provides an end-member test case suitable for assessing the accuracy and

264

robustness of numerical methods used to simulate real floods. The solution employs an

265

initial condition in which a triangular prism of static fluid is impounded by a dam face

266

normal to the slope, and the flood is triggered when the dam instantaneously vanishes.

267

Key aspects of the motion of the flood head and tail are illustrated by some elementary

268

features of our solution obtained directly from the untransformed shallow-water equations.

269

For example, the solution shows that the evolving speed of the flow front is the same as

270

that of a frictionless point mass with an initial velocity u


= 2 gH0 cos , where g is

271

the magnitude of gravitational acceleration, H0 is the initial height of water behind the

272

dam, and is the slope angle. Relative to motion of the flow front, motion of the tail is

273

delayed by a time proportional to cotan, because motion of the tail does begin until a

274

wave propagates upstream from the broken dam. This delay causes the downslope speed

275

of the tail to persistently lag behind that of the front, and as a consequence of this delay

276

and the fact that the tail subsequently accelerates like a frictionless point mass, the flood

277

wave elongates at a constant rate. Our solution describes evolution of the elongating

278

flood wave in terms of definite integrals that are readily evaluated using software such as

279

Mathematica. This evaluation shows that the flood wave is initially quite asymmetric but

280

becomes increasingly symmetric as time proceeds.

281

Finally, we note that extension of our solution to more complex dam-break flows involv-

282

ing materials other than ideal fluids may be possible. Motion of rock avalanches, snow

283

avalanches, and debris flows, for example, obeys equations that are mathematically simi-

284

lar to the shallow-water equations [Savage and Hutter , 1989; Pudasaini and Hutter , 2006;

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 18

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

285

Iverson and Denlinger , 2001; Mangeney-Castelnau et al., 2005; Balmforth and Kerswell ,

286

2005], and these phenomena are good candidates for further analytical study. In partic-

287

ular, the experimental and numerical results obtained by Greve et al. [1994] and Koch

288

et al. [1994] for dam-break avalanches of granular materials down steep chutes appear

289

very similar to results described in this paper.

290

Acknowledgments. We thank Professors J.M.N.T. Gray (University of Manchester)

291

and Randall LeVeque (University of Washington) for their helpful reviews of our work.

292

In particular, Professor LeVeque ran his shock-capturing numerical code Tsunamiclaw

293

(freely downloadable from http://www.amath.washington.edu/claw/) and found excel-

294

lent agreement with our solution. The work presented here was supported by the Swiss

295

National Science Foundation under grant number 200021-105193, the competence center

296

in Mobile Information and Communication Systems (a center supported by the Swiss Na-

297

tional Science Foundation under grant number 5005-67322), the Competence Center in

298

Environmental Sciences (TRAMM and APUNCH projects), and specific funds provided

299

by EPFL (vice-presidence `a la recherche). Use of trade names in this publication is for

300

information purposes only and does not constitute endorsement by the U.S. Geological

301

Survey.

D R A F T

September 26, 2007, 3:54pm

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

X - 19

Appendix A
302

In this paper, basal friction has been neglected. This assumption is likely to be valid

303

in the bulk of the flow since the bottom friction contribution is usually of low magnitude

304

compared to the inertia and pressure gradient terms in the momentum balance equation.

305

Close to the front, this assumption no longer holds because the flow depth drops to zero.

306

To estimate the typical extent of the friction-affected region, the usual approach is

307

to use a balance between friction and pressure gradient, i.e., if we use a Chezy law for

308

cos h/
x Cd
representing the bottom drag, we have g h
u2 , where Cd denotes a

309

Chezy-like coefficient, in the drag-affected region [Whitham, 1954; Hogg and Pritchard ,

310

and u are not explicitly known.


2004]. A difficulty arises here since h

311

and u for the head.


To proceed further in this analysis, we first need to approximate h

312

This can be readily done by making a first-order approximation of the integral represen-

313

tations (30) and (31) of t(r, s) and (r, s) for the head. Then solving the resulting linear

314

system to find r and s, we find


2 4
s= +
,
3 3t

(A1)

r = 2.

(A2)

315

Making use of Eq. (12) to find h and u and returning to dimensional variables, we finally

316

obtain

!2

1
xf x
=
h
,
9g cos
t

!
x
1
2 + uf ,
u =
3
t
317

318

(A3)
(A4)

where xf = 2 gH0 cos t + 21 gt2 sin denotes the front position and uf = 2 gH0 cos +
gt sin its velocity. A remarkable feature is that the flow-depth and velocity profiles in

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 20

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

319

the close vicinity of the front have exactly the same shape as those found for the Ritter

320

[1892] solution. Denoting = xf x, we find that within the tip region ( 0), the

321

dominant balance is
g cos

322

2 1
h
Cd u2f ,
t2

(A5)

which yields
3g

2
h
81Cd g cos2
u2f t3 .
t2

(A6)

323

At short times, uf 2 gH0 cos and therefore the extent of the drag-affected region

324

scales as t4/3
1/3

1/3

4Cd g 2/3 cos H0 t4/3 ,

(A7)

325

which is consistent with the scaling found for dam-break waves on horizontal planes

326

[Whitham, 1954; Hogg and Pritchard , 2004]. At long times, uf gt sin , which results

327

in a more pronounced dependence of on t


4Cd gt2 cos2/3 2.

(A8)

Appendix B
328

329

The Riemann function R can be computed as follows [Garabedian, 1964, see problem
9, 5.1, p. 150]. Let us consider a partial differential equation of the form
vxy +

330

1
(vx + vy ) = 0,
2x+y

(B1)

whose adjoint operator is


N [v] = 0, with N [v] = vxy (av)x (bv)y + cv, and a = b =

D R A F T

September 26, 2007, 3:54pm

1
,
2x+y

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED


331

and where c = 0. Following Garabedian [1964], we pose


v=

332

X - 21

(x + y)
(x )(y )
W (), with =
.
/2
/2
(x + ) (x + )
(x + )(y + )

We find that W satisfies the equation


2 W () + 4(1 ( + 1))W 0 () + (1 )W 00 () = 0,

333

whose solution is

"


W () = F
, , 1, ,
2 2
334

where F is the hypergeometric function. With = 3, x = r and y = s, we find the

335

solution to the adjoint problem (22) with N given by Eq. (19). Alternative representa-

336

tions (in particular, in terms of Legendre functions) can be obtained using properties of

337

F [Abramowitz and Stegun, 1964, see pp. 559562].

Appendix C
338

In this appendix we relate our results to those of Savage and Savage and Hutter [1989],

339

who obtained similarity solutions to the shallow-flow equations for motion of finite volumes

340

of frictional material down a uniform slope. Of particular relevance here is their parabolic

341

cap solution, which can be obtained by seeking symmetric flow-depth and velocity profiles

342

for the governing equations (8) and (9). In Fig. 5, we note that at sufficiently late times,

343

the flow-depth profile is bell-shaped, while the velocity profile is nearly linear with . This

344

prompts us to seek a solution, where the velocity profile is perfectly linear and takes the

345

value v = f at the front ( = f ), i.e.,


v(, t) =

D R A F T


f ,
f

September 26, 2007, 3:54pm

(C1)

D R A F T

X - 22

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

346

where f denotes the front position and f its velocity in the t plane. For the mo-

347

ment, f (t) is unknown; we expect that the similarity solution is the long-time asymptotic

348

solution of the boundary initial value problem solved above and therefore assume that

349

f 2t. Substituting v into the momentum balance equation (9), we derive an equation

350

for h
h
f
= ,

351

(C2)

whose integration provides


h(, t) =

1 f 2
( 2 ).
2 f f

(C3)

352

The flow-depth profile is parabolic and symmetric around = 0. Substituting the v and

353

h relations into the mass equation (8), we derive an equation for the front position f
d
(f f ) + f f = 0.
dt

354

Integrating this equation leads to the second-order differential equation


f2 f = c1 ,

355

(C4)

(C5)

with c1 a constant of integration, which can be determined using volume conservation


V=

Z f
f

2
2
h(, t)d = f2 f = c1 ,
3
3

(C6)

356

where V = 12 |xb | = 12 cotan is the initial volume of material. We can now find f from

357

(C5) using the boundary conditions


lim f = 2t and f (0) = 0.

(C7)

358

The former boundary condition enforces behavior similarity between this solution and the

359

one found above using the method of characteristics. The latter condition is somewhat
D R A F T

September 26, 2007, 3:54pm

D R A F T

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

X - 23

360

formal, but is consistent with our objective of finding the long-time asymptotic solution.

361

Integrating (C5) twice and using the boundary conditions (C7), we find an implicit relation

362

relating f to t
q

8 3V + 4 4 2 3 V

f
f
f
= 16t,
4 4f2 3f V + 3V ln

3V

(C8)

363

which is valid for > 3V/4. Differentiating this equation with respect to t, we find that

364

the front velocity is given by


q

f =

365

(C9)

We check that f 2 when f . The parabolic cap solution is given by


v(, t) =
h(, t) =

366

f (4f 3V)


f ,
f

(C10)

3V 2
2

,
f
4 f3

(C11)

with f given by (C8) and f given by (C9).

367

In Fig. 9, we have plotted the parabolic cap solution for t = 100. We also have also

368

shown the exact solution to the shallow-water equations. A key point is that although

369

both velocity profiles superimpose remarkably, there is a substantial difference in the

370

shape of the surge. For the exact solution, the flow-depth profile is always acute close to

371

the fronts since the flow-depth gradient drops to zero (see Appendix A), whereas for the

372

similarity solution, the height gradient at the front is nonzero (h/ = 3V/(2f )), which

373

results in a finite front angle that the remaining flow must accommodate.

References
374

375

Abramowitz, M., and I. A. Stegun (1964), Handbook of Mathematical Functions, National


Bureau of Standards, Washington.
D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 24
376

377

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

Ancey, C., S. Cochard, S. Wiederseiner, and M. Rentschler (2006), Front dynamics of


supercritical non-Boussinesq gravity currents, Water Resour. Res., 42, W08,424.

378

Ancey, C., S. Cochard, S. Wiederseiner, and M. Rentschler (2007), Existence and features

379

of similarity solutions for supercritical non-Boussinesq gravity currents, Physica D, 226,

380

3254.

381

382

Balmforth, N. J., and R. R. Kerswell (2005), Granular collapse in two dimensions, J. Fluid
Mech., 538, 399428.

383

Bouchut, F., A. Mangeney-Castelnau, B. Perthame, and J. P. Vilotte (2003), A new model

384

of Saint Venant and Savage-Hutter type for gravity driven shallow flows, C. R. Acad.

385

Sci. Paris ser. I, 336, 531536.

386

387

388

389

Carrier, G. F., and H. P. Greenspan (1958), Water waves of finite amplitude on a sloping
beach, J. Fluid Mech., 4, 97109.
Chanson, H. (2004), The Hydraulics of Open Channel Flow: An Introduction, 2nd ed.,
Elsevier Butterworth Heinemann, Amsterdam.

390

Chanson, H. (2006), Analytical solutions of laminar and turbulent dam break wave, in In-

391

ternational Conference on Fluvial Hydraulics River Flow 2006, vol. 1, edited by R. M. L.

392

Ferreira, E. C. T. L. Alves, J. G. A. B. Leal, and A. H. Cardoso, pp. 465474, Lisbon,

393

Portugal.

394

395

396

397

Costa, J. E. (1988), Floods from dam failure, in Flood Geomorphology, edited by V. R.


Baker, R. C. Kochel, and P. C. Patton, pp. 439461, Wiley, New York.
Courant, R., and K. O. Friedrich (1948), Supersonic Flow and Shock Waves, Intersciences
Publishers, New York.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 25

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED


398

Daly, E., and A. Porporato (2004a), Similarity solutions of nonlinear diffusion problems

399

related to the mathematical hydraulics and the Fokker-Planck equation, Phys. Rev. E,

400

70, 056,303.

401

402

403

404

405

406

407

408

409

410

Daly, E., and A. Porporato (2004b), A note on groundwater flow along a hillslope, Water
Resour. Res., 40, W01,601.
Dressler, L. (1952), Hydraulic resistance effect upon the dam-break functions, J. Res. Natl
Bur. Stand., 49 (3), 217225.
Dressler, R. F. (1978), New nonlinear shallow-flow equations with curvature, J. Hydraul.
Res., 16, 205222.
Dressler, R. H. (1958), Unsteady non-linear waves in sloping channels, Proc. R. Soc.
London ser. A, 247, 186198.
Fernandez-Feria, R. (2006), Dam-break flow for arbitrary slopes of the bottom, J. Eng.
Math., 54, 319331.

411

Garabedian, P. R. (1964), Partial Differential Equations, John Wiley & Sons, New York.

412

Greve, R., T. Koch, and K. Hutter (1994), Unconfined flow of granular avalanches along

413

414

415

416

417

418

419

a partly curved surface. I. Theory, Proc. R. Soc. London ser. A, 445, 399413.
Hogg, A. J. (2006), Lock-release gravity currents and dam-break flows, J. Fluid Mech.,
569, 6187.
Hogg, A. J., and D. Pritchard (2004), The effects of hydraulic resistance on dam-break
and other shallow inertial flows, J. Fluid Mech., 501, 179212.
Hunt, B. (1983), Perturbation solution for dam-break floods, J. Hydraul. Eng., 110, 1053
1071.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 26

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

420

Iverson, R. M., and R. P. Denlinger (2001), Flow of variably fluidized granular masses

421

across three-dimensional terrain. 1. Coulomb mixture theory, J. Geophys. Res., 106,

422

537552.

423

Karelsky, K. V., V. V. Papkov, A. S. Petrosyan, and D. V. Tsygankov (2000), The initial

424

discontinuity decay problem for shallow water equations on slopes, Phys. Lett. A, 271,

425

349357.

426

427

428

429

430

431

Keller, J. B. (2003), Shallow-water theory for arbitrary slopes of the bottom, J. Fluid
Mech., 489, 345348.
Kerswell, R. R. (2005), Dam break with Coulomb friction: a model for granular slumping?,
Phys. Fluids, 17, 057,101.
Kevorkian, J. (2000), Partial Differential Equations, Analytical Solution Techniques,
Springer, New York.

432

Koch, T., R. Greve, and K. Hutter (1994), Unconfined flow of granular avalanches along

433

a partly curved surface. II. Experiments and numerical computations, Proc. R. Soc.

434

London ser. A, 445, 415435.

435

436

Mangeney, A., P. Heinrich, and R. Roche (2000), Analytical solution for testing debris
avalanche numerical models, Pure Appl. Geophys., 157, 10811096.

437

Mangeney-Castelnau, A., B. Bouchut, J. P. Vilotte, E. Lajeunesse, A. Aubertin, and

438

M. Pirulli (2005), On the use of Saint-Venant equations for simulating the spreading of

439

a granular mass, J. Geophys. Res., 110, B09,103.

440

441

442

Peregrine, D. H., and S. M. Williams (2001), Swash overtopping a truncated plane beach,
J. Fluid Mech., 440, 391399.
Pudasaini, S. P., and K. Hutter (2006), Avalanche Dynamics, Springer, Berlin.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 27

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED


443

444

Ritter, A. (1892), Die Fortpflanzung der Wasserwellen, Zeit. Vereines Deutsch. Ing.,
36 (33), 947954.

445

Saint Venant, B. (1871), Theorie du mouvement non permanent des eaux, avec application

446

aux crues des rivi`eres et `a lintroduction des marees dans leur lit, C. R. Acad. Sci. Paris

447

ser. I, 173, 147154 237240.

448

449

450

451

Savage, S. B., and K. Hutter (1989), The motion of a finite mass of granular material
down a rough incline, J. Fluid Mech., 199, 177215.
Shen, M. C., and R. E. Meyer (1963), Climb of a bore on a beach. Part 3. Run-up, J.
Fluid Mech., 16, 113125.

452

Stoker, J. J. (1957), Water Waves, Interscience Publishers, New York.

453

Whitham, G. B. (1954), The effects of hydraulic resistance in the dam-break problem,

454

Proc. R. Soc. London ser. A, 227, 399407.

455

Whitham, G. B. (1974), Linear and Nonlinear Waves, John Wiley & Sons, New York.

456

Zauderer, E. (1983), Partial Differential Equations of Applied Mathematics, Pure and

457

458

459

Applied Mathematics, John Wiley & Sons, New York.


Zoppou, C., and S. Roberts (2003), Explicit schemes for dam-break simulations, J. Hydraul. Eng., 129, 1134.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 28

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

H0

Figure 1.

The initial configuration of the reservoir before the dam collapse.

5
C
4
3

F
2

1
0
0

10

x
Figure 2. Characteristics corresponding to the boundaries of the moving fluid volume.
Computation is for slope angle = /4.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 29

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

Table 1.

OF

Features of the boundaries delimiting the fluid domain.


c

t tan + 2

2t

OB 1 t tan /2
BC

Table 2.

t tan t2 tan /4 t 2(1 t tan ) 2

0
t tan 2

2t + cotan

Equations of the boundaries delimiting the fluid domain.


x

D R A F T

t range

OF

t2
2

tan + 2t

t0

OB

t2
4

tan t

0 t 2cotan

BC

t2
2

tan 2t + cotan

t 2cotan

September 26, 2007, 3:54pm

D R A F T

X - 30

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

+2

2
r
M(a, b)
b

B, C

Figure 3.

D R A F T

t=0

b
O
P
t = 1 r/2

Computation domain in the r s plane.

September 26, 2007, 3:54pm

D R A F T

X - 31

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

5
4

3
2
1

-2

x
Figure 4. Characteristics in the t plane for slope angle = /4. The r-characteristics
are shown as solid lines for r values ranging from 2 to 2, with an increment of 0.5. The
s-characteristics are shown as dashed lines for s values ranging from 2 to 2, with an
increment of 0.5.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 32

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

0.30

hH x,tL

0.25
0.20
0.15
0.10
0.05
0.00
-15

-10

-5

10

15

vH x,tL

1
0
-1
-2
-15

-10

-5

10

15

Figure 5.

Flow depth and velocity profiles in the t plane for slope angle = /4.

Profiles are shown for times t = 1, 2, 4, 8.

The dashed line represents the initial flow

depth (still water).

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 33

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

0.25

t=1

h( x,t )

0.2
0.15
t=2
0.1
t=4
0.05

t=8
0

10

20

30

40

x
10
t=8

u( x,t )

8
6
t=4
4
t=2
t=1

2
0
0

10

20

30

40

x
Figure 6.

Flow depth and velocity profiles in the x t plane for slope angle = /4.

Profiles are shown for times t = 1, 2, 4, 8. The dashed line represents the initial flow
depth.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 34

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

h( x,t )

0.8
0.6
0.4
0.2

-1 -0.5

0.5

1.5

2.5

x
Figure 7.

Flow depth profiles in the x t plane for slope angle = /4. Profiles are

shown for times t = 0.25, 0.5, 0.75,1. The dashed line represents the initial flow depth.

5
4

3
2
1

-2

x
Figure 8.

Characteristics in the x t plane for slope angle = /4.

The -

characteristics are shown as solid lines for values ranging from 2 to 2, with an increment
of 0.5. The -characteristics are shown as dashed lines for values ranging from 2 to 2,
with an increment of 0.5.

D R A F T

September 26, 2007, 3:54pm

D R A F T

X - 35

ANCEY ET AL.: DAM-BREAK PROBLEM ON SLOPING BED

2
0.0020

vH x,tL

hH x,tL

1
0.0015
0.0010

-1

0.0005
0.0000
-200

-100

100

200

-2
-200

-100

100

200

Figure 9. comparison between the exact solution to the shallow-water equations (solid
line) given implicitly by equations (30) and (31) and the parabolic cap solution (dashed
curve) given by equations (C10) and (C11) at t = 100. On the left: flow-depth profile; on
the right: flow-depth averaged velocity.

D R A F T

September 26, 2007, 3:54pm

D R A F T

Potrebbero piacerti anche