Dam-Break Modeling in Alpine Valleys

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

J. Mt. Sci. (2014) 11(6): 1429-1441 e-mail: [email protected] http://jms.imde.ac.

cn
DOI: 10.1007/s11629-014-3042-0

Dam-break Modeling in Alpine Valleys

Marco PILOTTI1*, Andrea MARANZONI2, Luca MILANESI1, Massimo TOMIROTTI1,


Giulia VALERIO1

1 DICATAM, Università degli Studi di Brescia, Via Branze 43, 25123 - Brescia, Italy
2 DICATeA, Parco Area delle Scienze 181/A, 43124 - Parma, Italy

*Corresponding author, e-mail: [email protected]; 2nd author, e-mail: [email protected];


3rd author e-mail: [email protected]; 4th author, e-mail: [email protected];
5th author, e-mail: [email protected]

Citation: Pilotti M, Maranzoni A, Milanesi L, et al. (2014) Dam-break modeling in alpine valleys. Journal of Mountain
Science 11(6). DOI: 10.1007/s11629-014-3042-0

© Science Press and Institute of Mountain Hazards and Environment, CAS and Springer-Verlag Berlin Heidelberg 2014

Abstract: Dam-break analysis is of great importance equations; Alpine valleys; Case study; Hydraulic
in mountain environment, especially where reservoirs hazard
are located upstream of densely populated areas and
hydraulic hazard should be assessed for land planning
purposes. Accordingly, there is a need to identify Introduction
suitable operative tools which may differ from the
ones used in flat flood-prone areas. This paper shows
The risk assessment of dam failure is a topic
the results provided by a 1D and a 2D model based on
the Shallow Water Equations (SWE) for dam-break that deserves great attention since, although its
wave propagation in alpine regions. The 1D model probability of occurrence is rather small, the
takes advantage of a topographic toolkit that includes related consequences might be catastrophic. The
an algorithm for pre-processing the Digital Elevation problem could become even worse in the future
Model (DEM) and of a novel criterion for the because of the intensification of extreme
automatic cross-section space refinement. The 2D meteorological events, the general increase of
model is FLO-2D, a commercial software widely used exposure, and the ageing of the structures. The
for flood routing in mountain areas. In order to verify growth of exposure is related to the increased
the predictive effectiveness of these numerical models, number of new reservoirs built over the last
the test case of the Cancano dam-break has been decades in developing countries and to the
recovered from the historical study of De Marchi
expansion of strongly populated areas in developed
(1945), which provides a unique laboratory data set
nations. Therefore, in many countries, technical
concerning the consequences of the potential collapse
of the former Cancano dam (Northern Italy). The guidelines concerning dam safety require hydraulic
measured discharge hydrograph at the dam also studies aimed to estimate hazard levels in dam-
provides the data to test a simplified method recently break flood prone areas. This requires the use of
proposed for the characterization of the hydrograph suitable mathematical tools capable of effectively
following a sudden dam-break. modeling the dynamics of very rapid floods.
In alpine valleys, dam-break wave propagation
Keywords: Dam-break modeling; Shallow water may be very complex because of the interaction of
the impulsive nature of the process with the strong
Received: 1 March 2014 irregularity of natural topographies. This particular
Accepted: 16 June 2014

1429
J. Mt. Sci. (2014) 11(6): 1429-1441

feature suggests the adoption of purposely devised propagation in flood plains where the one-
computational tools. Despite the hypothesis of dimensionality of the flow, the uniformity of
spatially gradually varied flow is often violated in velocity, and horizontality of water level across the
practical applications, especially in mountain cross-section are violated. Accordingly, a
environment, SWE are usually accepted in the comparative analysis of the performances provided
literature and used in the engineering practice by 1D and 2D modelling is useful, although rather
dealing with flood propagation problems (e.g., rare in literature (e.g., Horritt and Bates 2002;
Soares Frazão et al. 2003; Capart et al. 2003; Tayefi et al. 2007).
Aureli et al. 2008a; Petaccia et al. 2008; Pilotti et Pilotti et al. (2011) showed that for an accurate
al. 2011). Indeed, SWE conveniently combine description of very steep and irregular natural
computational efficiency and satisfactory bathymetries by means of a 1D approach, it could
reconstruction of the physical phenomenon. be necessary to have closely spaced cross-sections,
This paper compares a one-dimensional (1D) which would be impossible to acquire by
and a two-dimensional (2D) shallow water model conventional field survey procedures. In this
for dam-break wave propagation in alpine regions. direction, the increasing availability of detailed
The 1D de Saint-Venant equations are solved using DEMs has fostered the development of efficient
a finite volume numerical scheme based on the geographical information systems (GIS) able to
approach proposed by Capart et al. (2003). It takes analyse, process, and manage massive quantities of
advantage of a topographic toolkit that includes an spatial data. Many algorithms for the automatic
algorithm for pre-processing the Digital Elevation extraction of drainage networks from digital
Model (DEM) and a novel criterion for the elevation models have been proposed in literature
automatic cross-section refinement. The 2D model (e.g., Pilotti et al. 1996), almost exclusively for
is based on the commercial software FLO-2D, space-distributed hydrological and environmental
widely used for hydraulic hazard mapping in applications (e.g., Pilotti and Bacchi 1997). In this
mountain areas, mostly for debris-flow and paper these methods are used to deal with a
mudflow. Recently its performance in granular hydraulic problem and an algorithm that performs
debris-flow propagation was comparatively tested a space adaptive sampling along the drainage
by Wu et al. (2013). Accordingly, it is interesting to network obtained from a DEM is presented.
explore its performances also when dam-break Consequently, the cross-sections (and the
wave propagation is concerned. associated geometric information) were extracted
Although 2D models are gaining increasing with an adaptive spatial resolution.
popularity and provide a more accurate description Commonly, the reliability of the results
of the hydrodynamic process, the 1D obtained using SWE is firstly verified on the basis
schematization is still widely applied whenever the of theoretical test cases with analytical solution
topography of the valley and the discharge involved that usually are characterized by simple geometry
justify this assumption. A priori, the main reasons and mostly by frictionless bottom (Ritter 1892;
are the greater practitioners’ acquaintance with 1D Stoker 1948; Thacker 1981; McDonald et al. 1997;
modelling, the more limited computational burden Liska and Wendroff 1998; Liska and Wendroff
and implementation effort, and the need of a 1999).
sufficiently detailed DEM for 2D studies, which However, the numerical modeling of flood
may be difficult to obtain in some cases. Moreover, wave propagation in mountain regions demands
1D models could be still competitive against 2D that the chosen mathematical and numerical tool
models with respect to the capability of predicting simultaneously deals with several stumbling blocks
the flood dynamics and the inundation extent in (e.g. wet/dry boundaries, non prismatic geometry,
rivers and streams (Horritt and Bates 2002). In non uniform bed slope, transcritical flows, shocks)
very wide domains, in order to limit the that in the simplified literature test cases are dealt
computational time, 1D and 2D approaches can be with separately. For this reason, laboratory
sinergically coupled (Fernández-Nieto et al. 2010): measures and field data concerning real events are
the former can be used for the wave propagation in valuable information for validating numerical
channel-like stretches, the second for the flood modeling, even if they do not provide the exactness

1430
J. Mt. Sci. (2014) 11(6): 1429-1441

of analytical solutions. Data collected during for the characterization of the hydrograph
laboratory investigations usually concern idealized following a sudden dam-break.
situations and schematic geometries (e.g., Soares
Frazão 2007; Soares Frazão and Zech 2007; Aureli
et al. 2008b). Very few are the experimental data 1 1D Modeling
from physical models reproducing a real
topography. In particular, the most known physical One-dimensional flood propagation modeling
model is probably the one built by Electricité de requires a proper treatment of the source term,
France in 1964 to reproduce the Malpasset dam- especially in the presence of natural complex
break (e.g., Goutal 1999; Hervouet and Petitjean bathymetries. In fact, in the classical formulation of
1999; Valiani et al. 2002), where maximum water the SWE (e.g. Toro 2001), the geometrical source
depths were measured at 14 gauges. Other terms involving the spatial gradients of bottom
experimental data widely used in literature (Soares elevation and of cross-sectional width cannot be
Frazão and Testa 1999; Brufau et al. 2002; Caleffi easily evaluated from the topographical
et al. 2003) concern the flood wave propagating in information, usually based on a limited number of
the physical model built by ENEL–HYDRO (Italy) cross-sections. In order to cope with these
reproducing a portion of the Toce River valley difficulties, Capart et al. (2003) proposed an
(located in the Northern Italian Alps) and approximate evaluation of the pressure force acting
including water depth hydrographs measured at 33 on the surface of the wetted boundary that allows
selected points. On the other hand, documented transferring the geometric source terms (that
historical dam-break events are rather rare. Among account for non-prismaticity and bed slope) within
these, the well-known catastrophic events of the momentum flux, enhancing mass and
Malpasset (Goutal 1999), Tous (Alcrudo and Mulet momentum conservation properties of the
2007), St. Francis (Begnudelli and Sanders 2007), numerical model in the simulations over real
Gleno (Pilotti et al. 2011), and Lawn Lake dam bathymetries. As shown by Capart et al. (2003) and
(Jarrett and Costa 1986) can be mentioned. In Chen et al. (2007), the consequent reformulation of
general, the available information is rather limited, the governing equations does not introduce
scattered and uncertain, but nonetheless it can be remarkable inaccuracies, even in the presence of
useful to define dam-break test cases and verify the hydraulic jumps or shock waves.
capability of numerical models to reproduce the The numerical integration of the modified
aspects of the flood event that are most relevant for SWE was performed by the first-order algorithm of
hazard classification (maximum water levels and Capart et al. (2003), which makes use of the PFP
velocity, flood extent, and timing of the flood wave). (Pavia Flux Predictor) upwind method proposed by
This paper shows for the first time the Braschi and Gallati (1992) for the numerical flux
application of the numerical models to the case prediction. A detailed description of this numerical
study examined by De Marchi (1945) and model can be found in Pilotti et al. (2011). However,
concerning the propagation along the Adda River despite its robustness and suitability to unsteady
(Northern Italy) of the flood wave consequent to river flow modelling over natural topography, this
the potential collapse of the first Cancano dam as a scheme requires a very fine computational mesh in
possible war target during World War II. The the presence of a steep water surface slope
resulting report of De Marchi is very interesting, combined with a steep bathymetry, typical of
since it synergistically mixes theoretical, mountain streams (Capart et al. 2004). This kind
experimental, and numerical considerations. In of spatial discretization can increase the
particular, the measured discharge hydrographs computational effort.
reported in that paper can be used to assess the
performances of SWE and the reliability of 1.1 DEM pre-processing and cross-section
numerical results in mountain applications. extraction
Furthermore, the measured discharge hydrograph
at the dam provides the opportunity to validate the In river flow computations, an extensive
simplified method proposed by Pilotti et al. (2010) numerical DEM pre-processing is usually needed

1431
J. Mt. Sci. (2014) 11(6): 1429-1441

to extract the geometry of the modelled system. If a support area. According to the first method, a
1D approach is adopted, a topographic drainage line is considered as a channel if the
representation of the valley by cross-sections must upstream drained area A is larger than a suitable
be provided. This can be done using the channel threshold At. Due to the dependence of the bed
and hillslope drainage networks obtained from the shear stress on the local gradient, the second
informative content of the source DEM. Here, a method takes into account also the local slope
specific GIS tool performing this task was used (for through the condition: A>AtSα, S and α being the
details, see Pilotti et al. 1996). local slope and an exponent, respectively.
First, the pre-processing algorithm removes Figures 1a-1d show four different channel
depressions (where flow-lines converge) and flat networks derived for an upper reach of the Adda
regions (where data points are at the same River valley, where the Cancano dam is located.
elevation), that actually would prevent the Both drainage and channel networks are encoded
identification of the connected basin drainage using a tree-like pointers structure, that logically
network. Using the enhanced topographic reproduces the network topology allowing an
information, a space filling drainage network is efficient manipulation of its informative content.
then derived following the steepest descent After that the location of the upstream section has
directions. Whilst the space filling drainage been selected on the channel network, the code
network reproduces all the flow paths inside the extracts the hydraulic path down to the basin outlet
basin, in river flow modelling the channel network along with the information on local elevations,
(that is a subset of the overall drainage network) slopes, and distances. The river reach is then
must be identified. With this aim, the algorithm discretized according to the original DEM
here employed filters the space filling drainage resolution Δl: accordingly, the reach length
network either according to the method based on between two consecutive points is Δl or 20.5 Δl. At
the fixed threshold contributing area or to the each point along the channel the local flow
method based on the slope dependent critical direction is identified drawing the secant line

Figure 1 Different channel networks obtained from the DEM of the alpine watershed considered in the paper using
the slope dependent algorithm (A>At Sα, S being the local bottom slope). (a) At=0.1 km2, α =0; (b) At=0.1 km2, α=1.7;
(c) At=0.2 km2, α=2; (d) At=0.5 km2, α=1.7. The white area in (d) is the location of Cancano reservoir, whereas the
black spot is the outlet of the watershed at Ponte Cepina.

1432
J. Mt. Sci. (2014) 11(6): 1429-1441

connecting the second


point upstream and the
second point downstream.
However, the automatic
recognition of the flow
direction is not
straightforward when the
stream bends or
meanders. Finally, the
DEM surface is sampled
orthogonally to the local
flow direction by bilinear
interpolation in order to
extract cross-section
profiles and derive the
geometric information
required for the 1D
hydrodynamic simulation.
(a) (b)
Figure 2 Transcritical steady flow through a sinusoidal contraction-expansion: (a)
1.2 Spatial comparison between numerical results obtained by different mesh sizes; (b) plan view
discretization of the channel and distribution of the εa parameter for different space discretizations.
h and discharge Q higher than 20%.
Although the use of 1D SWE is a standard in
We propose to evaluate the acceptability of the
literature, both in steady and unsteady flows, the
space discretization on the basis of the comparison,
problem of the definition of appropriate criteria for
at a fixed stage h , between the uniform flow stage-
cross-section selection and mesh refinement has
discharge relationships of two consecutive cross-
apparently been overlooked in the past and few
sections spaced ∆x. Accordingly, a relative
literature contributions deal with it (e.g., Samuels
deviation
1990; Castellarin et al. 2009). This topic is here
briefly analyzed and a novel method for automatic Q(x + Δx ,h ) − Q(x ,h )
ε(x ,h ,Δx ) =
cross-section space refinement is suggested. Q(x ,h ) (2)
As an example, let us consider the simple case is defined. In the discussed test case, introducing a
of a steady flow through a sequence of a first order accurate approximation, this coefficient
contraction-expansion in a frictionless, horizontal, reads:
60 m long, rectangular channel where the width b
varies according to the equation: 1 db( x )  5 2 
ε( x, h , Δx ) =  −  Δx .
3 dx  b( x ) b( x ) + 2h 
 (3)
  x − 30 
 5 − 1.5 1 + cos  2π  for x − 30 ≤ 15 This quantity can be averaged in each cross-
b( x) =    30  .
 5 section between 0 and H, being H a reasonably
otherwise
expected maximum value for the local water depth
(1) during the flood:
The inflow discharge is set to 20 m3/s
and a H

 ε( x, h , Δx) dh . (4)
transmissive boundary condition is assigned 1
ε a ( x , Δx ) =
downstream. This test problem was numerically H
solved for different mesh sizes ∆x using the 0

numerical model described above. The results The maximum value of ε a ( x , Δ x ) along the
shown in Figure 2a emphasize a considerable discretized channel can be assumed as an indicator
dependence on cross-section spacing: a value of ∆x of the appropriateness of the selected space
greater than 10 m implies maximum error on stage discretization (see Figure 2b).

1433
J. Mt. Sci. (2014) 11(6): 1429-1441

Real situations are by far more complex than form, so that the numerical scheme is not ideal to
the one shown in this example, since steep slopes accurately reproduce shock-waves or hydraulic
and continuous changes in cross-section shape jumps.
could occur in addition to changes of width.
However, the indicator εa takes into account also
these effects, providing a simple criterion for the 3 The Cancano Dam Test Case
automatic local refinement of the cross-sections
along the channel. Therefore, after the selection of The effectiveness of the operative tools
the channel from the DEM and the automatic described in the previous sections was verified on
extraction of the cross-sections (along with the basis of the Cancano dam-break test case
geometric quantities and uniform flow stage- recovered from the historical paper of De Marchi
discharge relationships), the local value εa (x, ∆x) is (1945).
computed. According to this vector, cross-sections In 1943, after the bombing of several dams in
are locally and automatically refined where the Ruhr region, the possibility that the Cancano
necessary, obtaining an improved and adaptive 1D dam (see Figure 3 and Figure 4) could become a
mesh.

2 2D Modeling

The 2D numerical model used in this paper is


the commercial code FLO-2D (2009). This choice
was suggested by the compliance of this software
with the requirements of FEMA about flood hazard
studies, with particular reference to steep
mountain alluvial fans, where mud and debris
flows can be expected. With reference to this type
of events, FLO-2D was compared in the past with
other 2D codes for debris flow simulation
(Rickenmann et al. 2006; Wu et al. 2013). However,
FLO-2D is a general code suitable also for flood
propagation. Accordingly, verification of its
effectiveness in clear water dam-break cases can be
of interest.
FLO-2D solves the full 2D SWE on a Cartesian
grid using an explicit, central, finite difference
numerical scheme along eight potential flow
directions. Each velocity computation is essentially
one-dimensional and is performed independently
for each direction. The Courant-Friedrichs-Lewy
condition and the stability criterion developed by
Ponce and Theurer (1982) for nonlinear equations
are applied, as well as a check on the percentage
time change in the flow depth within each
computational cell. If any of these stability criteria
is not met, the time step is reduced and the
computations are repeated, so that the method can
require lengthy computer runs to simulate steep
rising flood waves. According to O’Brien et al. Figure 3 Location of the Cancano dam in the Adda
(1993), the solved equations are not in conservative River valley (Northern Italian Alps).

1434
J. Mt. Sci. (2014) 11(6): 1429-1441

Figure 4 Historical picture and topography of the former Cancano dam viewed from downstream and in plan (from
ANIDEL 1953).
war target was considered. The dam was a concrete box with prototypal dimension of approximately
gravity structure and was built in the period 1220 × 704 m; a gap of 265 m along its shortest
between 1924 and 1929 for hydroelectric purposes. side represented the Cancano dam. The sudden
It was 60 m high and 265 m long, and retained and complete removal of the plate covering the gap
24 × 106 m3 of water. Nowadays it is submerged allowed to simulate the collapse of the dam.
into the new Cancano reservoir that was built in Both total and partial collapses of the dam
the 50’s with a total storage capacity of about were considered. In both cases, discharge
124 × 106 m3. hydrographs were measured at three locations: a
De Marchi was commissioned to evaluate the section just downstream of the dam, at Section 23
expected hydraulic consequences of the possible and at Ponte Cepina. Each discharge hydrograph
collapse of the Cancano dam and the resulting was obtained by graphic derivation from the filling
study provides an outstanding example of coupling time series recorded by using a floating device in a
between numerical and experimental analysis. calibrated tank. For each collapse scenario, several
According to the fact that “it is practically runs were carried out in the same test conditions to
impossible to apply the usual equations of verify the consistency of the measuring procedure.
Hydraulics to the mostly irregular, sometimes The final hydrographs were obtained by averaging.
extremely irregular, bottom of mountain valleys” Furthermore, the delimitation of the flooded areas
(translation from De Marchi 1945), for the first between Section 23 and Ponte Cepina was
rugged reach downstream of the reservoir an
experimental approach was adopted. Actually,
along the 8 km reach from the Cancano dam down
to the village of Premadio (Section 23 in Figure 5),
the Adda River valley has a mean bed slope of 7.5%
with long stretches characterized by bed slope
greater than 20%. From Section 23 to Ponte Cepina
(see Figure 5) the valley is more regular and mean
bed slope reduces to about 1%. In this stretch De
Marchi applied also a simplified numerical
approach based on SWE.
A 1:500 scale physical model reproducing the
bathymetry of the 16 km long valley stretch from
the Cancano dam to Ponte Cepina was built
according to the Froude similitude. As one can
Figure 5 Longitudinal profile of the thalweg of the
deduce from the original paper, the Cancano lake Adda River between the former Cancano dam and
was reproduced as a horizontal rectangular-shaped Ponte Cepina.

1435
J. Mt. Sci. (2014) 11(6): 1429-1441

accomplished. These data provide a unique mass conservation law can be used (Pilotti et al.
benchmark for validating numerical models for 2013),
flood routing in mountain environment. Qp
Q(t ) = (6)
3
 Qp

3.1 Reconstruction of the hydrograph at the 1 + 2V
t
dam  
where V is the water volume initially stored in the
De Marchi’s experimental data also allows the reservoir. Figure 6 shows an excellent match
validation of a procedure recently proposed by between the experimental and theoretical
Pilotti et al. (2010) for the computation of the hydrographs.
hydrograph following a partial dam-break.
However, this procedure is not suitable to the
reconstruction of the whole hydrograph in the case
of a horizontal reservoir like the one built by De
Marchi; accordingly, here it will be used to
compute the peak discharge only.
Considering the overall volume stored in the
reservoir model and its area, the initial water depth
during the tests can be computed as 0.056 m, that
corresponds to h0=27.9 m in the prototype. For this
depth and for the total collapse of the dam, the
Ritter formula provides a peak discharge of
36327 m3/s. Due to the relative dimension of the
dam with respect to the length of the rectangular-
Figure 6 Comparison between measured and
shaped box where it was opened, the sudden and calculated discharge hydrograph at the dam section.
complete removal of the dam has a similar effect of
a partial dam-break with a breach ratio
4 Results and Discussion
b/B=265/740= 0.376. Pilotti et al. (2010) have
shown that, whenever a partial dam-break occurs,
the widening of the negative wave front upstream The dam-break wave following the total
of the breach causes an amplification of the peak collapse of Cancano dam was simulated using the
discharge with respect to the Ritter value. Pilotti et two numerical models presented in Sections 1 and
al. (2010) provides an amplification factor for the 2. Only a 20 × 20 m resolution DEM is available for
Ritter discharge as a function of the breach ratio this part of the valley and it was used as a basis for
(see Eq. 12 of the cited paper), that in the case of the description of the computational domain. We
rectangular cross-section can be rewritten as: completed the DEM by including the valley stretch
b (now submerged by the reservoir) between the old
Qp  27  B and the present Cancano dam. To this purpose,
  = k (b / B ) (5) elevation contour lines obtained from historical
Bh0 gh0  8 
official topographic maps were used. Unfortunately,
where k(b/B) can be interpolated by the tabulated the original cross-sections used by De Marchi are
values in the cited paper and is equal to 0.2412 in not available and the cross-section profiles for the
this case. Accordingly, the computed peak 1D model were extracted from the DEM using the
discharge Qp is 49690 m3/s, in excellent agreement semi-automatic procedure described in Section 1.
with the value of approximately 50000 m3/s In both models, the discharge hydrograph
measured by De Marchi. De Marchi timed the peak measured at the dam cross-section (see Figure 6)
discharge immediately after the removal of the was imposed as upstream boundary condition,
dam and measured a monotonically decreasing whereas a free outflow condition was applied at the
hydrograph. In order to interpret its falling limb, outlet. Moreover, a dry bed condition was initially
the curve obtained by coupling a weir flow with the assumed along the modeled reach. For a proper

1436
J. Mt. Sci. (2014) 11(6): 1429-1441

numerical treatment of wetting and drying fronts, a hand, peak discharge seems much less affected by
1 mm and 3 cm water depth threshold was imposed grid size.
in the 1D and 2D models, respectively, accepting a The sensitivity of numerical results with
maximum final mass error less than 1% with respect to n was also analyzed. In Figure 8 the gray
respect to the total volume initially stored in the bands represent the envelope of the discharge
reservoir. The Courant-Friedrichs-Lewy number hydrographs predicted by the 1D model for values
was set at 0.8. of the Manning coefficient within the ranges
De Marchi made the concrete surface of the mentioned above. One can notice that this
physical model artificially rough to reproduce the parameter has a stronger influence on the timing of
natural irregularity of the upper part of the valley the flood than on the peak discharge value. As
and the effect of vegetation. Although an expected, the wetting front arrival time decreases
estimation of the roughness of the physical model with the reduction of the Manning coefficient.
is missing in the historical paper, the morphology Moreover, the effect of roughness on the arrival
of the valley upstream of Section 23 (very steep and time is less evident in the upper reach of the valley
winding) suggests the adoption of a Manning’s
coefficient n between 0.1 and 0.067 m−1/3s, as
confirmed in literature studies concerning
mountain creeks (e.g. Jarrett 1984). Similarly,
visual inspection of the downstream reach between
Section 23 and Ponte Cepina may justify values
between 0.05 and 0.033 m−1/3s. This choice is
supported by the fact that a similar value (n =
0.05 m−1/3s) was adopted by De Marchi for the
numerical modeling of the dam-break propagation
downstream of Ponte Cepina.
Operating with the 1D model, first a sensitivity
analysis on mesh resolution was accomplished. A
non uniform mesh (N = 125 cross-sections
characterized by a mean spacing Δxm of about Figure 7 Sensitivity of the 1D model to spatial
130 m) was initially adopted, doubling the number resolution: computed discharge hydrographs at two
of cross-sections used in the experimental study. selected cross-sections based on n=0.067m−1/3s
upstream of Section 23 and n=0.04m−1/3s
Then spatial resolution was gradually increased downstream.
obtaining three refined meshes with N = 667
(Δxm=24.4 m), 1332 (Δxm=12.2 m) and 4987 cross-
sections (Δxm=3.3 m) respectively. Figure 7 shows
the discharge hydrographs computed at the
locations indicated in Figure 5 (Section 23 and
Ponte Cepina) using different meshes. The
Manning’s coefficient was set at n= 0.067 m−1/3s for
the reach upstream of Section 23 and at n=
0.04 m−1/3s for the downstream stretch of the valley.
The numerical scheme achieves convergence for
Δxm =12.2 m. Consequently, the corresponding
mesh was chosen for the calculations. Furthermore,
flood arrival time is sensitive to mesh resolution: in
particular, as one could expect, mesh refinement
induces a delay of the arrival time of the wetting
Figure 8 Sensitivity of the 1D model to roughness
front. Actually, a coarse mesh smooths bed coefficient: envelope of the computed discharge time
irregularities and, consequently, induces a faster series at two selected cross-sections based on the mesh
propagation of the dam-break wave. On the other characterized by Δxm=12.2 m.

1437
J. Mt. Sci. (2014) 11(6): 1429-1441

that is characterized by very steep bed slopes. On Adda River and its tributary upstream of Section
the whole, the speed of the flood wave seems well 23, where strong two-dimensional effects occur.
reproduced by the numerical model. Figure 8 also Other discrepancies can be observed in the
highlights that the 1D model systematically floodplain downstream of Section 23, where
overestimates both the peak discharge and the flooded areas are considerably overestimated, as
steepness of the rising limb of the discharge one could expect. As one can see in Figure 10, the
hydrographs. This behavior is probably due to the 2D simulation reproduces more faithfully the
difficulty of the scheme in capturing two- observed inundation extent. The agreement
dimensional expansions and related attenuation between observed and computed flooded areas can
effects at lateral confluences. The falling limb is be quantified by using the performance index
less influenced by roughness and shows a good suggested by Horritt and Bates (2002):
agreement with the experimental one. Figure 9
Asim ∩ Aobs
shows the hydrographs calculated by FLO-2D. The P =100 ⋅ (7)
Asim ∪ Aobs
results show a minor sensitivity to the Manning
coefficient in the same range explored with the 1D where A makes reference to the simulated and
model. In both sections the peak discharge is well observed extension of flooded areas. The P index
reproduced, although there is a delay in the arrival ranges from 100 (perfect agreement) down to 0 in
time of the wave and the computed rising limb is
much steeper than the observed one.

Figure 9 Sensitivity of the FLO-2D model to the


Manning coefficient. Case (a) n= 0.10 m−1/3s in the
reach upstream of Section 23 and n= 0.04 m−1/3s for
the downstream part of the domain. Case (b) n= 0.067
m−1/3s and n= 0.03 m−1/3s, respectively.

Finally, the calculated inundation maps were


compared with the experimental results of De
Marchi (Figure 10). The inundated areas were
delimited on the basis of the envelope of maximum
water depths calculated by the 1D numerical model
according to the assumption of horizontality of
water level in the cross-sections. This mapping was
repeated for different roughness values, showing
that the shorelines of flooded areas are almost
insensitive to the Manning coefficient. The 1D
Figure 10 Comparison between calculated and
model does not faithfully match the observed measured flooded areas between Section 23 and Ponte
pattern, especially at the confluence between the Cepina.

1438
J. Mt. Sci. (2014) 11(6): 1429-1441

the case of complete uncorrelation between to overestimate the extent of flooding. Accordingly,
predicted and observed inundated areas. In the whilst the use of a 2D approach throughout the
case of the simulation accomplished by the 1D domain would imply a considerable increase of the
model, P is 63.8 while using FLO-2D the P index computational time (non compatible with a real-
rises up to 73.4. In considering these values, one time application in the examined case), a 2D
should take into account that De Marchi description of the flood limited to these areas could
represented this part of the domain by using 32 greatly improve the computational efficiency.
cross-sections: accordingly, one can expect slight Accordingly, the 1D modelling of wave propagation
variations of the bathymetry which could affect in channel-like stretches can provide a boundary
flood propagation in the shallowest parts of the condition for 2D flood propagation in flood plains.
flow field. When a 1D approach is used in steep and
From the computational point of view, the run irregular alpine valleys, the problem of cross-
of the 1D code lasts approximately 0.7 hour. The sections extraction could not be solved using
same simulation accomplished with FLO-2D traditional surveying techniques. In the considered
(calibrated to obtain the results here presented) case, the distance between two consecutive cross-
takes approximately 72 hours using i7-core PC with sections cannot exceed 20 m on average and the
4 GB RAM. Other 2D solvers can certainly be more convergence of the solution is obtained for an
efficient that the one currently implemented in FLO- average grid spacing of approximately 12 m. In order
2D but there is a clear disproportion between the to solve this problem, in this paper we outline a
computational time required by the two approaches. novel procedure to derive cross-sections by
Finally, one can note that although a long stretch of automatic sampling of the DEM channel network.
the river downstream of Section 23 is not strictly This can be derived by the analysis of the steepest
suitable to a 1D simulation, also in this area 1D terrain directions, as usually done in hydrologic
results are conservative without being too unrealistic. rainfall-runoff modeling. The introduction of a
Given that the peak time is of minor interest in such control criterion based on the comparison of
a fast-evolving process, the predicted peak discharge consecutive normal flow depths allowed an
and extent of flooded areas are sufficiently close to automated mesh refinement where required. In this
the measured ones to retain engineering interest. way, an accurate and efficient 1D reconstruction of
However, a 1D approach only provides the average the topography was performed.
velocity and depth cross-sections and, accordingly, The Cancano test case, which was recovered
is not suitable to the use of physically-based from the historical paper of De Marchi, is an
criterions for hazard mapping (Xia et al. 2014; extraordinary benchmark for the validation of
Milanesi et al. 2014) that require a local description numerical SWE solvers in alpine valleys. The three
of the flow field. measured hydrographs and the measured extent of
flooding documented in the original paper were
used to compare the results provided by the two
5 Conclusions mathematical models and to show that SWE
provide a powerful prediction tool even in the
The study of dam break in alpine areas mountain environment, where the underlying
involves several stumbling blocks and can be hypotheses are far from being rigorously satisfied.
tackled using 1D and 2D models for flood The measured hydrograph at the dam was also
propagation. In this paper we compared the results used to verify the predictive effectiveness of a
provided by a 1D code (Pilotti et al. 2011) and the simple theoretical method to evaluate the dam-
2D code FLO-2D. The 1D approach is break peak discharge.
computationally less demanding and simpler to
implement and is acceptable whenever the extent
of the discharge and the narrowness of the valley Acknowledgements
justify the underlying hypothesis. In wide domains,
this approach could be effectively used to route the We thank the anonymous reviewers for their
flood wave as far as floodplain areas, where it tends valuable suggestions that contributed to the

1439
J. Mt. Sci. (2014) 11(6): 1429-1441

improvement of the paper. The study has been (Grant agreement 265280).
developed within the European Project Kulturisk

References

Alcrudo F, Mulet J (2007) Description of the Tous dam break Hervouet JM, Petitjean A (1999) Malpasset dam-break revisited
case study (Spain). Journal of Hydraulic Research 45 (Extra with two-dimensional computations. Journal of Hydraulic
Issue): 45-57. DOI: 10.1080/00221686.2007.9521832 Research 37(6): 777-788. DOI: 10.1080/00221689909498511
ANIDEL, The Federation of the Italian Electric Power Horritt MS, Bates PD (2002) Evaluation of 1D and 2D
Companies [Associazione Nazionale Imprese Produttrici e numerical models for predicting river flood inundation.
Distributrici di Energia Elettrica] (1953) The dams of the Journal of Hydrology 268(1-4): 87-99. DOI: 10.1016/S0022-
Italian hydroelectric plants [Le dighe di ritenuta degli 1694(02)00121-X
impianti idroelettrici italiani]. Vol. 7, ANIDEL, Rome, Italy Jarrett RD (1984) Hydraulics of high gradient streams. Journal
(In Italian). of Hydraulic Engineering 110(11): 1519-1539. DOI: 10.1080/
Aureli F, Maranzoni A, Mignosa P, et al. (2008a). 2D numerical 00221689909498511
modeling for hydraulic hazard assessment: a dam-break case Jarrett RD, Costa JE (1986) Hydrology, geomorphology, and
study. Proceedings of River Flow 2008. Çeşme, Turkey. pp dam-break modeling of the July 15, 1982, Lawn Lake Dam
729-736. and Cascade Lake Dam failures, Larimer County, Colorado.
Aureli F, Maranzoni A, Mignosa P, et al. (2008b) Dam-break US Government Printing Office, Washington, USA.
flows: acquisition of experimental data through an imaging Liska R, Wendroff B (1998) Composite schemes for
technique and 2D numerical modeling. Journal of Hydraulic conservation laws. SIAM Journal on Numerical Analysis
Engineering 134(8): 1089-1101. DOI: 10.1061/(ASCE)0733- 35(6): 2250-2271. DOI: 10.1137/S0036142996310976
9429(2008)134:8(1089) Liska R, Wendroff B (1999) Two-dimensional shallow water
Begnudelli L, Sanders BF (2007) Simulation of the St. Francis equations by composite schemes. International Journal for
dam-break flood. Journal of Engineering Mechanics, 133(11): Numerical Methods in Fluids 30(4): 461-479. DOI:
1200-1212. DOI: 10.1061/(ASCE)0733-9399(2007)133:11(1200) 10.1002/(SICI)1097-0363(19990630)30:4<461::AID-FLD850
Braschi G, Gallati M (1992) Conservative flux prediction >3.0.CO;2-4
algorithm for the explicit computation of transcritical flow in MacDonald I, Baines M, Nichols N, et al. (1997) Analytic
natural streams. Proceedings of Hydrosoft ‘92. Southampton, benchmark solutions for open-channel flows. Journal of
England. pp 381-395. Hydraulic Engineering 123(11): 1041-1045. DOI: 10.1061/
Brufau P, Vázquez-Cendón ME, García-Navarro P (2002) A (ASCE)0733-9429(1997)123:11(1041)
numerical model for the flooding and drying of irregular Milanesi L, Pilotti M, Ranzi R, et al. (2014) Methodologies for
domains. International Journal for Numerical Methods in hydraulic hazard mapping in alluvial fan areas. In: Evolving
Fluids 39(3): 247-275. DOI: 10.1002/fld.285 Water Resources Systems: Understanding, Predicting and
Caleffi V, Valiani A, Zanni A (2003) Finite volume method for Managing Water–Society Interactions, Proceedings of
simulating extreme flood events in natural channels. Journal ICWRS2014, Bologna, Italy, 4-6 June 2014, IAHS Publication
of Hydraulic Research 41(2): 167-177. DOI: 10.1080/0022168 n.364, Wallingford (UK), ISBN: 978-1-907161-42-1, 267-272.
0309499959 O’Brien J, Julien P, Fullerton W (1993) Two-dimensional water
Capart H, Eldho TI, Huang SY, et al. (2003) Treatment of flood and mudflow simulation. Journal of Hydraulic
natural geometry in finite volume river flow computations. Engineering, 119(2): 244-261. DOI: 10.1061/(ASCE)0733-
Journal of Hydraulic Engineering 129(5): 385-393. DOI: 9429(1993)119:2(244)
10.1061/(ASCE)0733-9429(2003)129:5(385) Petaccia G, Natale L, Savi F (2008) Simulation of Sella Zerbino
Capart H, Eldho TI, Huang SY, et al. (2004) Closure to catastrophic dam-break. Proceedings of River Flow 2008.
“Treatment of natural geometry in finite volume river flow Çeşme, Turkey. pp 601-607.
computations”. Journal of Hydraulic Engineering 130(10): Pilotti M, Gandolfi C, Bischetti GB (1996) Identification and
1048-1049. DOI: 10.1061/(ASCE)0733-9429(2004)130:10 analysis of natural channel networks from digital elevation
(1048.2) models. Earth Surface Processes and Landforms 21(11): 1007-
Castellarin A, Di Baldassarre G, Bates PD, et al. (2009) Optimal 1020. DOI: 10.1002/(SICI)1096-9837(199611)21:11<1007::
cross-sectional spacing in Preissmann scheme 1D AID-ESP704>3.0.CO;2-V
hydrodynamic models. Journal of Hydraulic Engineering Pilotti M, Bacchi B (1997) Distributed evaluation of the
135(2): 96-105. DOI: 10.1061/(ASCE)0733-9429(2009)135: contribution of soil erosion to the sediment yield from a
2(96) watershed. Earth Surface Processes and Landforms 22(13):
Chen J, Steffler PM, Hicks FE (2007) Conservative formulation 1239-1251. DOI: 10.1002/(SICI)1096-9837(199724)22:13<
for natural open channels and finite-element implementation. 1239::AID-ESP839>3.0.CO;2-K
Journal of Hydraulic Engineering 133(9): 1064-1073. DOI: Pilotti M, Tomirotti M, Valerio G, Bacchi B (2010) Simplified
10.1061/(ASCE)0733-9429(2007)133:9(1064) method for the characterization of the hydrograph following a
De Marchi G (1945) On the dam-break wave following the sudden partial dam break. Journal of Hydraulic Engineering
collapse of the Cancano dam [Sull’onda di piena che 136(10): 693-704. DOI: 10.1061/(ASCE)HY.1943-7900.000
seguirebbe al crollo della diga di Cancano]. L’Energia Elettrica 0231
22: 319-340. (In Italian) Pilotti M, Maranzoni A, Tomirotti M, Valerio G (2011) 1923
Fernández-Nieto ED, Marin J, Monnier J (2010) Coupling Gleno dam-break: case study and numerical modeling.
superposed 1D and 2D shallow-water models: source terms in Journal of Hydraulic Engineering 137(4): 480-492. DOI:
finite volume schemes. Computers & Fluids 39(6): 1070-1082. 10.1061/(ASCE)HY.1943-7900.0000327
DOI: 10.1016/j.compfluid.2010.01.016 Pilotti M, Tomirotti M, Valerio G, et al. (2013) Discussion of
FLO-2D Software, Inc. (2009) FLO-2D Reference manual "Experimental investigation of reservoir geometry effect on
(Version 2009), Arizona, USA. dam-break flow, by A. Feizi Khankandi, A. Tahershamsi and S.
Goutal N (1999) The Malpasset dam failure. An overview and Soares-Frazão. Journal of Hydraulic Research 51(2): 220-222.
test case definition. Proceedings of 4th CADAM meeting. DOI: 10.1080/00221686.2013.765165
Zaragoza, Spain. Ponce VM, Theurer FD (1982) Accuracy criteria in diffusion

1440
J. Mt. Sci. (2014) 11(6): 1429-1441

routing. Journal of the Hydraulics Division 108(6): 747-757. 2007.9521830


Rickenmann D, Laigle D, McArdell BW, et al. (2006) Stoker JJ (1948) The formation of breakers and bores.
Comparison of 2D debris-flow simulation models with field Communications on Pure and Applied Mathematics 1:1-87.
events. Computational Geosciences 10(2): 241-264. DOI: Tayefi V, Lane SN, Hardy RJ, et al. (2007) A comparison of one-
10.1007/s10596-005-9021-3 and two-dimensional approaches to modelling flood
Ritter A (1892) The propagation of water waves [Die inundation over complex upland floodplains. Hydrological
Fortpflanzung der Wasserwellen]. Zeitschrift des Vereines Processes 21(23): 3190-3202. DOI: 10.1002/hyp.6523
Deutscher Ingenieure 36(3): 947-954 (In German). Thacker WC (1981) Some exact solutions to the nonlinear
Samuels PG (1990) Cross section location in one-dimensional shallow-water wave equations. Journal of Fluid Mechanics
models. In: White WR (Ed.), International Conference on 107: 499-508. D OI: 10.1017/S0022112081001882
River Flood Hydraulics. Wiley, Chichester, England. pp 339- Toro EF (2001) Shock-capturing Methods for Free-Surface
350. Shallow Flows. John Wiley & Sons Ltd., Chichester, England.
Soares Frazão S, Testa G (1999) The Toce River test case: Valiani A, Caleffi V, Zanni A (2002) Case study: Malpasset dam-
numerical results analysis. Proceedings of the 3rd CADAM break simulation using a two-dimensional finite volume
workshop. Milan, Italy. method. Journal of Hydraulic Engineering 128(5): 385-393.
Soares Frazão S, Zech Y, et al. (2003) IMPACT, Investigation of DOI: 10.1061/(ASCE)0733-9429(2002)128:5(460)
extreme flood processes and uncertainty. 3rd Project Wu YH, Liu KF, Chen YC (2013) Comparison between FLO-2D
Workshop Proceedings. Louvain-la-Neuve, Belgium. and Debris-2D on the application of assessment of granular
Soares Frazão S (2007). Experiments of dam-break wave over a debris flow hazards with case study. Journal of Mountain
triangular bottom sill. Journal of Hydraulic Research 45 Science 10(2): 293-304. DOI:10.1007/s11629-013-2511-1
(Special Issue): 19-26. DOI: 10.1080/00221686.2007. Xia J, Falconer RA, Wang Y, et al. (2014). New criterion for the
9521829 stability of a human body in floodwaters. Journal of Hydraulic
Soares Frazão S, Zech Y (2007). Experimental study of dam- Research 52(1): 93-104. DOI: 10.1080/00221686.2013.875
break flow against an isolated obstacle. Journal of Hydraulic 073
Research 45 (Special Issue): 27-36. DOI: 10.1080/00221686.

1441

You might also like