0% found this document useful (0 votes)
36 views

Three-Dimensional Numerical Modelling of Real-Field Dam-Break Flows: Review and Recent Advances

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
36 views

Three-Dimensional Numerical Modelling of Real-Field Dam-Break Flows: Review and Recent Advances

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 26

water

Review
Three-Dimensional Numerical Modelling of Real-Field
Dam-Break Flows: Review and Recent Advances
Andrea Maranzoni 1, * and Massimo Tomirotti 2

1 Department of Engineering and Architecture, University of Parma, 43124 Parma, Italy


2 Department of Civil, Environmental, Architectural Engineering and Mathematics, University of Brescia,
25121 Brescia, Italy; [email protected]
* Correspondence: [email protected]

Abstract: Numerical modelling is a valuable and effective tool for predicting the dynamics of the
inundation caused by the failure of a dam or dyke, thereby assisting in mapping the areas potentially
subject to flooding and evaluating the associated flood hazard. This paper systematically reviews
literature studies adopting three-dimensional hydrodynamic models for the simulation of large-
scale dam-break flooding on irregular real-world topography. Governing equations and numerical
methods are analysed, as well as recent advances in numerical techniques, modelling accuracy, and
computational efficiency. The dam-break case studies used for model validation are highlighted. The
advantages and limitations of the three-dimensional dam-break models are compared with those of
the commonly used two-dimensional depth-averaged ones. This review mainly aims at informing
researchers and modellers interested in numerical modelling of dam-break flow over real-world
topography on recent advances and developments in three-dimensional hydrodynamic models so
that they can better direct their future research. Practitioners can find in this review an overview of
available three-dimensional codes (research, commercial, freeware, and open-source) and indications
for choosing the most suitable numerical method for the application of interest.

Keywords: dam break; flooding; numerical modelling; real-world topography; review; three-
dimensional models

Citation: Maranzoni, A.; Tomirotti,


M. Three-Dimensional Numerical 1. Introduction
Modelling of Real-Field Dam-Break Dam-break floods are caused by the uncontrolled release of water stored in a reservoir
Flows: Review and Recent Advances. due to a total or partial collapse of a constructed or natural dam. Such phenomena can
Water 2023, 15, 3130. https://
result in potentially catastrophic consequences and many more casualties than other kinds
doi.org/10.3390/w15173130
of floods and disasters (e.g., [1,2]). Zhang et al. [3] documented 1443 failures of constructed
Academic Editor: Anargiros I. Delis dams and 1044 failures of landslide dams worldwide over the past two centuries and pro-
vided a list of the 20 most significant dam failures (each causing more than 500 casualties),
Received: 22 June 2023
resulting in more than 44,000 fatalities. Among these 20 selected dam disasters, the most
Revised: 25 August 2023
catastrophic one is the 1975 Banqiao dam failure (China), which led to the inundation of
Accepted: 27 August 2023
an area of approximately 12,000 km2 and the loss of more than 26,000 lives, followed by
Published: 31 August 2023
the disasters of the Vajont dam (Italy) in 1963, the South Fork dam (US) in 1889, and the
Machhu-II dam (India) in 1979, each with more than 2000 casualties [1,3,4]. In addition,
dam-break flooding can cause huge economic losses and extensive environmental dam-
Copyright: © 2023 by the authors. age [5]. For example, the 1976 Teton dam disaster (US) resulted in only 11 fatalities, but the
Licensee MDPI, Basel, Switzerland. flood covered an area of 77 km2 and reached up to 250 km downstream of the dam, causing
This article is an open access article more than USD 400 million of damage [6].
distributed under the terms and Accordingly, the assessment of the flooding hazard associated with hypothetical dam-
conditions of the Creative Commons break scenarios has gained great importance and attracted considerable attention in the last
Attribution (CC BY) license (https:// decades, both in engineering practice (e.g., [7,8]) and research (e.g., [9–12]), for dam-break
creativecommons.org/licenses/by/ flood risk management, emergency response, and flood hazard mitigation planning.
4.0/).

Water 2023, 15, 3130. https://doi.org/10.3390/w15173130 https://www.mdpi.com/journal/water


Water 2023, 15, x FOR PEER REVIEW 2 of 29

Water 2023, 15, 3130


dam-break flood risk management, emergency response, and flood hazard mitigation 2 of 26
planning.
In principle, a dam-break problem can be studied using either an experimental [13]
or a mathematical modelling approach [14,15], or both together (e.g., [16]). However, nu-
In principle, a dam-break problem can be studied using either an experimental [13]
merical modelling has become attractive and increasingly popular due to its flexibility and
or a mathematical modelling approach [14,15], or both together (e.g., [16]). However,
cost-effectiveness and the continuous growth of computing capacities, which allows the
numerical modelling has become attractive and increasingly popular due to its flexibility
processing of a large amount of data and solving mathematical models of great complex-
and cost-effectiveness and the continuous growth of computing capacities, which allows the
ity and significant predictive capability [17]. Moreover, the current availability of several
processing of a large amount of data and solving mathematical models of great complexity
commercial and freeware Computational Fluid Dynamics (CFD) software codes,
and significant predictive capability [17]. Moreover, the current availability of several
equipped with friendly interfaces and including effective pre- and post-processing tools,
commercial and freeware Computational Fluid Dynamics (CFD) software codes, equipped
facilitates the diffusion of numerical models in flood hazard analyses [18]. The high-cost
with friendly interfaces and including effective pre- and post-processing tools, facilitates
and time-consuming complex operations required to carry out accurate laboratory exper-
the diffusion of numerical models in flood hazard analyses [18]. The high-cost and time-
imentation have determined that the physical modelling approach is currently mainly
consuming complex operations required to carry out accurate laboratory experimentation
adopted
have to obtainthat
determined experimental
the physical data useful forapproach
modelling the validation of numerical
is currently mainly models
adopted[13,14].
to
Usually, dam-break flood hazard assessment and
obtain experimental data useful for the validation of numerical models [13,14]. mapping are performed using
depth-averaged
Usually, dam-breaktwo-dimensional
flood hazard (2D) models, and
assessment which solve the
mapping are 2D shallowusing
performed waterdepth-
equa-
tions (SWEs)
averaged through finite(2D)
two-dimensional difference,
models, finite
whichvolume,
solve the or 2Dfinite element
shallow watermethods
equations(e.g.,
[15,19–21]) to predict relevant hydraulic quantities associated with flood
(SWEs) through finite difference, finite volume, or finite element methods (e.g., [15,19–21]) hazards (namely,
tomaximum flood depth
predict relevant and velocity).
hydraulic quantities Despite the simplifying
associated assumptions
with flood hazard (namely, underlying
maximum the
2D SWEs, there are many computational challenges in numerically
flood depth and velocity). Despite the simplifying assumptions underlying the 2D SWEs, solving these equa-
tions,are
there such
many as shock-capturing
computational challengescapability,intreatment
numerically of wet
solvingandthese
dry fronts,
equations, treatment
such asof
bottom and friction
shock-capturing sourcetreatment
capability, terms, reproduction
of wet and dry of fronts,
flow regime
treatmenttransitions,
of bottomand andpreserva-
friction
tion orterms,
source achievement
reproductionof stationary
of flow or steady-flow
regime conditions.
transitions, In any case,
and preservation orthe SWEs strictly
achievement of
apply to flows
stationary with no significant
or steady-flow conditions.curvature of the
In any case, thefree
SWEs surface and
strictly negligible
apply to flows vertical
with no ac-
celeration (and hence with nearly hydrostatic pressure distribution)
significant curvature of the free surface and negligible vertical acceleration (and hence with over small bottom
slopeshydrostatic
nearly (e.g., [21,22]). However,
pressure in gravity-driven
distribution) over small geophysical
bottom slopes flows,
(e.g., the terrain
[21,22]). may be
However,
invery steep, at least
gravity-driven locally, especially
geophysical flows, thein mountain
terrain may be regions or near
very steep, topographic
at least singulari-
locally, especially
inties [23–25].regions
mountain Moreover, or nearthetopographic
hydrostatic singularities
pressure assumption is violated
[23–25]. Moreover, theinhydrostatic
curvilinear
flows, which
pressure can also
assumption is occur
violated as in
a result of dam
curvilinear failures,
flows, which especially
can also during
occur asthe first stages
a result of damof
the motion
failures, (e.g., [26,27])
especially during or theinfirst
case the flooding
stages wave propagates
of the motion (e.g., [26,27])inorthein presence of bends
case the flooding
(e.g., propagates
wave [28]), contractions (e.g., [29,30]),
in the presence bottom
of bends singularities
(e.g., (e.g., [31,32]),
[28]), contractions or obstacles
(e.g., [29,30]), bottomand
structures (e.g.,
singularities (e.g.,[33–35]).
[31,32]),For example,and
or obstacles Figure 1 shows(e.g.,
structures two [33–35]).
pictures of Forthe impact of
example, a dam-
Figure 1
break two
shows wave against
pictures of athe
prismatic
impact ofblock [34]. Such
a dam-break a physical
wave against aprocess
prismatic is characterised
block [34]. Such by
amarked
physicalthree-dimensional
process is characterised by marked three-dimensional (3D) features.
(3D) features.

Figure 1. Three-dimensional features of a dam-break wave impacting a structure. The pictures were
Figure 1. Three-dimensional features of a dam-break wave impacting a structure. The pictures were
taken during the laboratory investigation performed by Aureli et al. [34]. Time t commences with
taken during the laboratory investigation performed by Aureli et al. [34]. Time t starts from the
the sudden gate removal.
sudden gate removal.
Local 3D effects cannot be reproduced by the 2D SWEs [36], with consequent possible
Local 3D effects cannot be reproduced by the 2D SWEs [36], with consequent possible
limitations in the predictive capability of 2D shallow-water models and inaccuracies in
limitations in the predictive capability of 2D shallow-water models and inaccuracies in the
the prediction of the relevant hydraulic variables, such as flood inundation extent, maxi-
prediction of the relevant hydraulic variables, such as flood inundation extent, maximum
mum flood depths, and impact loads on structures [37]. More general formulations of the
flood depths, and impact loads on structures [37]. More general formulations of the 2D
depth-averaged SWEs have been proposed in the literature to overcome this drawback
without resorting to more computationally expensive 3D models. In these enhanced formu-
lations, some of the restrictive assumptions of the classic shallow-water model are relaxed
Water 2023, 15, 3130 3 of 26

while retaining its robustness and simplicity. For example, steep-slope SWEs (SSSWEs)
were introduced to simulate shallow flows over steep terrain (e.g., [24,25,38]). Boussinesq-
type models (e.g., [23,39,40]), as well as vertically averaged and moment (VAM) equations
(e.g., [41]) or depth-averaged equations incorporating an “enhanced” gravity (e.g., [42,43]),
can be used to simulate non-hydrostatic flows, preserving the vertical momentum balance
and including the effect of the vertical flow acceleration.
Recent advances in computing performance have fostered the application of 3D nu-
merical models [44], which offer an improved predictive capability [45] and a more accurate
description of the flow features, especially where vertical flow acceleration cannot be ne-
glected and the pressure distribution is far from hydrostatic [19]. Three-dimensional models
with different degrees of complexity (and, consequently, different computational costs)
can be used, namely [40] direct numerical simulation (DNS) models, which numerically
solve the Navier–Stokes equations (NSEs), resolving all turbulence spatial and temporal
scales; large-eddy simulation (LES) models, which solve the filtered NSEs, thus ignoring
the smallest length scales; and models that solve the Reynolds-averaged Navier–Stokes
equations (RANS), coupled with a closure turbulence model [46]. DNS and LES models
are significantly more computationally expensive than the RANS ones, especially at the
high Reynolds numbers encountered in environmental free-surface flows. Therefore, the
application of DNS and LES models in this field is feasible only in limited domains [19].
Consequently, RANS solvers are the most common choice for large-scale real-world dam-
break flood simulations. However, since the flow depth does not explicitly appear in
the basic equations of 3D models, great care has to be devoted to spatial discretization
in the vertical direction, and additional computational effort is required for free surface
tracking [37]. Several free-surface-tracking techniques have been proposed in the literature.
The best known are the Volume of Fluid (VOF; [47]) and the Level Set [48] methods (or a
combination of the two in coupled Level Set–VOF methods; e.g., [49]). The VOF model
is the most popular in dam-break flow simulations [50]. An alternative strategy to solve
the 3D governing equations, avoiding the construction of a computational mesh and the
adoption of complex free-surface-tracking algorithms, is based on meshless particle-based
methods, such as Smoothed Particle Hydrodynamics (SPH; [51]), which has also been
applied in the modelling of dam-break flow (e.g., [52,53]), even on real-world topography,
and other environmental applications [54,55].
However, the high computational cost of 3D models is still a significant limitation,
especially in large-scale field studies or when high spatial accuracy is required. This
limitation has hindered the diffusion of 3D models in the past for real-field applications,
favouring the 2D depth-averaged ones. Hence, high-performance computing, such as
parallel or Graphics Processing Unit (GPU) computing, is today a valuable support to
enhance the computational efficiency of 3D models and reduce computational time.
Dam-break models (both 2D and 3D) are usually validated against experimental
data of laboratory test cases [13], which, however, are typically schematic (e.g., [44,56])
or include isolated singularities (e.g., [50,57–59]). Instead, real dam-break events involve
irregular topography and constitute far more challenging benchmarks for numerical models.
However, well-documented historical dam-break events are scarce and often characterised
by uncertain available information [14], making their use arduous for complete validation
of the numerical models (especially for 3D ones, in which several parameters are involved).
Nonetheless, they provide valuable information about the different degrees of reliability of
the models in reproducing dam-break flow features over real-world topography.
Table 1 provides a comparative summary of the advantages and shortcomings of 3D
and depth-averaged 2D models. Given the increasing diffusion of 3D CFD models and
their improved predictive capabilities compared with the 2D ones, this review focuses on
3D modelling of large-scale real-field dam-break floods.
Water 2023, 15, x FOR PEER REVIEW 4 of 29

Water 2023, 15, 3130 4 of 26

their improved predictive capabilities compared with the 2D ones, this review focuses on
3D modelling of large-scale real-field dam-break floods.
Table 1. Summary and comparison of the advantages and shortcomings of 3D and 2D models.
Table 1. Summary and comparison of the advantages and shortcomings of 3D and 2D models.
Model Advantages Shortcomings
Model Advantages Shortcomings
Easy to build and implement Limitations due to the
Easy to build and implement cheap
Computationally shallow-water
2D Limitations due assumptions
to the
Computationally
Few cheap to calibrate
parameters (hydrostatic distribution of
2D
depth-averaged Few parameters to calibrate shallow-water assumptions
depth-averaged (roughness) pressure
(hydrostatic distribution of pressure
(roughness)
Robust and stable and small bottom slopes)
and small bottom slopes)
Robust and stable Laborious to build and
High accuracyHigh accuracy Laborious to build and implement
implement
(mildassumptions)
(mild restrictive restrictive assumptions) Complex calculations
3D 3D Complex calculations
Reproduction of
Reproduction of non-hydrostatic ef- Computationally
Computationallyexpensive
expensive
non-hydrostatic effects
fects Several
Several parameters
parameters involved
involved

Figure
Figure 22 shows
showsaasketch
sketchofofthethemethodological
methodological framework
framework of this study.
of this The The
study. system-
sys-
atic review is restricted to publications focusing on 3D CFD modelling
tematic review is restricted to publications focusing on 3D CFD modelling of dam-break of dam-break flow
and
flowproviding
and providing application examples
application of numerical
examples simulation
of numerical of dam-break
simulation floodingflood-
of dam-break over
irregular real-world topography. A comparative analysis is performed
ing over irregular real-world topography. A comparative analysis is performed on the on the documents
collected,
documents and the information
collected, relevant to 3D
and the information CFD modelling
relevant to 3D CFD is modelling
reported inistabular
reportedformin
(Section 2). The results of the analysis allow statistical information on
tabular form (Section 2). The results of the analysis allow statistical information on keykey items of the
reviewed
items of the documents
reviewed(such as year (such
documents of publication,
as year ofsoftware status,
publication, model status,
software type, andmodelnu-
merical
type, and scheme)
numericalto bescheme)
obtained. to The discussionThe
be obtained. of the review findings
discussion is centred
of the review on the
findings is
following
centred onaspects: improvements
the following in simulation in
aspects: improvements accuracy,
simulationmodel validation
accuracy, model and calibra-
validation
tion, improvements
and calibration, in computational
improvements efficiency, and
in computational improvements
efficiency, in result visualization
and improvements in result
(Section 3). Finally,
visualization conclusions
(Section 3). Finally,are drawn, outlining
conclusions implications
are drawn, outliningthat can facilitate
implications thatprac-
can
tical applications and future research on 3D CFD models for the simulation
facilitate practical applications and future research on 3D CFD models for the simulation of of large-scale
dam-break flooding onflooding
large-scale dam-break irregularonreal-world topographytopography
irregular real-world (Section 4).(Section 4).

Figure 2. Schematic of the methodological framework of this review study.


Figure 2. Schematic of the methodological framework of this review study.
1.1.
1.1. Motivations
Motivations ofof the
the Present
Present Review
Review
There is a vast literature overviewing, describing,
There is a vast literature overviewing, describing, comparing,
comparing, and and evaluating
evaluating floodflood
inundation
inundation numerical
numerical modelsmodels for
for flood
flood risk
risk assessment
assessment and and water
water resources
resources management
management
For instance,
(e.g., [15,18,19]). For instance, Teng
Tengetetal.al.[15]
[15]conducted
conducteda acomprehensive
comprehensive review
review of of dif-
differ-
ferent modelling
ent modelling approaches
approaches (empirical,
(empirical, hydrodynamic,
hydrodynamic, andand simplified
simplified conceptual
conceptual mod-
models),
els), highlighting
highlighting their their advantages
advantages and drawbacks
and drawbacks and discussing
and discussing the sources
the sources of uncer-
of uncertainty in
tainty in flood inundation modelling. Mudashiru et al. [18] provided
flood inundation modelling. Mudashiru et al. [18] provided a similar overview, focusing a similar overview,
focusing on floodmapping.
on flood hazard hazard mapping.
Bates [19]Bates [19] surveyed
surveyed recent advances
recent advances in floodplain
in floodplain inundation in-
undation
modelling, modelling,
discussing discussing theofability
the ability of the numerical
the numerical modelsmodels to reproduce
to reproduce the phys-
the physical as-
ical aspects
pects of theofflooding
the flooding process.
process. LuoLuo et al.
et al. [60]
[60] andand Mignotand
Mignot andDewals
Dewals[61] [61] limited
limited the
area of focus, reviewing specifically urban flood simulation models and analysing their
hydrologic and hydrodynamic component modules. More More recently, Kumar et al. [62] and
Avila-Aceves et al. [63] offered offered a broad overview of different
different modelling approaches and
numerical techniques for simulating large-scale inundations, providing insight into their
strengths and weaknesses.
Only asome
few ofofthe thepreviously
previously mentioned
mentioned reviews
reviewstouch on the the
mention use use
of 3Dofhydrodynamic
3D hydrody-
modelsmodels
namic for floodforinundation simulations
flood inundation [15,19]. However,
simulations to our knowledge,
[15,19]. However, no literature
to our knowledge, no
reviews exist on the specific topic of 3D hydrodynamic modelling of large-scale dam-break
flows on real-world topography. Recent advances in 3D dam-break modelling, the growing
Water 2023, 15, 3130 5 of 26

attention given to 3D CFD models in recent times due to enhanced computational efficiency,
and the prospect of using 3D models extensively in engineering practice in the near future
for dam-break flood risk assessment are the main motivations for this review.

1.2. Objectives of the Present Review


This review aims to inform researchers new to the field and more experienced re-
searchers of the recent advances in 3D CFD dam-break modelling on real-world topography,
thereby enabling them to have a general, updated overview of this topic. The comparative
analysis of the relevant references selected can help modellers, practitioners, emergency
response agencies, and dam owners remain up to date on the latest developments in
dam-break simulation tools and choose the most suitable numerical model for the specific
application of interest. Finally, this review study provides state-of-the-art information to
direct future research on 3D dam-break numerical modelling for real field applications.

2. Review and Comparative Analysis


A systematic and careful search was performed in the most well-known scientific
databases (based on the keywords: “dam-break”, “3D numerical modelling”, “real topogra-
phy”, “complex topography”, and similar) to ensure wide coverage of the existing literature
on 3D numerical modelling of large-scale dam-break flows over real-world topography.
Mainly the academic literature (i.e., journal or conference articles) was considered; the grey
literature was not included.
The documents retrieved in the literature survey are arranged in Table 2, including
studies on dam spillway flows or overtopping flows reporting the numerical simulation of
the subsequent flood propagation downstream. Table entries are organised in chronological
order. References that contain multiple real-world case studies are repeated for each of
these and appear in different rows of the table (i.e., one row for each case study).
Table 2 provides the following, most significant information.
(1) Relevant references retrieved. Multiple references are reported in the same table
row when details providing a complete description of the case study analysed can be
obtained from various articles. In the case of duplicate studies, the most complete one was
considered, and subordinately, the earliest one.
(2) Code/software name, if available. Besides well-known commercial (e.g., TELEMAC-
3D [64]; FLOW-3D [65]) or open-source CFD software codes (e.g., OpenFOAM [66]; Dual-
SPHsysics [67]) widely used in many hydrodynamic applications, numerical codes (some-
times with no name) developed by universities or research centres for research purposes
(e.g., [68]) appear in the table.
(3) Basic equations of the simulation models: in order of increasing complexity, the Euler
equations, the Navier–Stokes equations, and the RANS equations for incompressible flow.
In the last case, a closure turbulence model is coupled with the governing equations. Lattice
Boltzmann methods (based on the Boltzmann equation and simulating the flow through
collision models of fictitious particles moving on a discrete lattice grid) are used more
rarely in dam-break flooding simulations. The table also reports the rheological model
coupled with the hydrodynamic equations in mudflow modelling of tailings dam breaks
and the erosional model used in the analysis of geomorphic dam-break flows over an
erodible bottom.
(4) Numerical methods used for spatial and temporal discretization of the model equa-
tions. The indication of the numerical scheme used for the simulations is accompanied by
the specification of the discretization technique (finite volume, finite element, SPH, etc.).
(5) Dam-break case studies considered to demonstrate the applicability of the numerical
models. Some case studies concern historical dam-break events, while others concern
hypothetical dam failures. The former case studies, if well documented, can be used for
model validation and to evaluate the model’s ability to reproduce real-field dam-break
flooding. Case studies of the latter type concern model applications to real-world situations,
mainly aimed at assessing dam-break flood hazards in potentially floodable areas.
Water 2023, 15, 3130 6 of 26

(6) Details about the computational domain and the spatial resolution adopted in the numer-
ical simulation of the case studies considered (i.e., the number of grid cells in mesh-based
models or fluid particles in meshless models). The mesh type (structured or unstructured)
and the shape of the grid elements are specified for mesh-based models.
(7) Outcomes of the numerical modelling. Typically, model results include flooded areas,
flood depth contour maps and flow velocity fields at selected times, contour maps of the
maximum values of flood depth and velocity magnitude, time series of flood depth and
flow velocity at given locations, and time series of flow discharge at selected cross-sections.
(8) Focus of the studies. The focus may be on the model validation, the prediction
or reconstruction of the inundation dynamics, or the 3D effects due to flow curvature.
Different aspects are sometimes examined contextually.
(9) Computational efficiency. The simulation (physical) time and the corresponding
computational run-time are reported for the case studies (when these data are available),
along with strategies implemented to improve computational efficiency.
(10) Publication year. This bibliographic information is useful to place each contribution
in time, outlining the evolution of the numerical models over time and research trends. If
multiple references are associated with a single table row, the publication year of the oldest
one is reported.
(11) Status of the CFD model. This can be commercial, freeware, open-source, or research.
Water 2023, 15, 3130 7 of 26

Table 2. Overview of studies on 3D hydrodynamic modelling of large-scale dam-break flooding over real-world topography.

(6) (9)
(4) (8)
(1) (2) (3) (5) Computational (7) Computational (10) (11)
Reference Model Name Model Type 1 Numerical Focus of
Method 2 Case Study Domain and Output Data Year Status
Elements
the Study Efficiency 3

Historical 1963
overtopping of the
Vajont dam (Italy) Modelled area extent:
Volume of the N/A Performance of the
rockslide: (the reservoir and Water surface numerical Simulation time:
Navier–Stokes and 270 million m3 the Vajont River at selected times; transverse technique 220 s
Roubtsova and Weakly Stored water
N/A continuity downstream) water surface profiles at a Reconstruction Run time/ 2006 Research
Kahawita [69] * compressible SPH volume: Number of particles: of the event
equations selected cross-section in simulation time:
115 million m3 Comparison with ~74
N/A the reservoir
(reservoir water
level provided) Particle spacing: field observations
Overtopping N/A
water volume:
30 million m3
Flooding dynamics
Flow fields (velocity for different
Modelled area extent: collapse scenarios
the reservoir and a valley magnitude) and flooded areas
Historical Comparison with
stretch downstream of the at selected times; motion of Simulation time:
1928 St. Francis wall fragments; flow field data
Cleary et al. [70]; Navier–Stokes and Weakly dam 25 min
dam break Modelling of the 2010
N/A continuity (California) Number of particles: discharge hydrograph at Run time/ Research
Prakash et al. [71] equations compressible SPH motion of dam simulation time:
Water volume: the dam site; flood arrival
1.4 × 106 times and maximum flood
wall blocks N/A
47 million m3 Particle spacing: 3D effects
depths at selected Sensitivity on
4m
locations particle resolution
(4 m; 6 m; 8 m)
Modelled area extent: Flow fields (velocity
the reservoir and a valley magnitude) and flooded areas Flooding dynamics
Hypothetical stretch downstream of the at selected times 3D effects
Geheyan dam break dam (different views); Simulation time:
Cleary et al. [70]; Effect of different
Navier–Stokes and Weakly (China) Number of particles: discharge hydrograph at the 60 min
N/A continuity dam failure Run time/ 2010 Research
Ye et al. [72]; Water volume: 1.3 × 106 (fluid), scenarios
compressible SPH dam site;
Cleary et al. [73] equations 3.12 billion m3 flow discharge hydrographs Modelling of the
simulation time:
Different dam 1.9 × 106 (boundaries) N/A
failure scenarios Particle spacing: at selected sections; flood motion of dam
15 m (fluid), depth hydrographs at wall blocks
30 m (boundaries) selected locations
Modelled area extent:
The reservoir (assumed to be
of prismatic shape) and
~250 m-long valley reach
downstream of the dam
(according to a 1:20 scale Qualitative
Weakly Ski-jump spillway reconstruction of Simulation time:
Navier–Stokes and physical model) 16 s
compressible and of the Number of particles: Spillway flow dynamics; the
Lee et al. [74] * N/A continuity spillway process Run time/ 2010 Research
equations truly Goulours dam 9.366 × 105 flooded areas at selected times simulation time:
incompressible SPH (France) Spillway flow ~2.7 × 104
(wall particles:
features
2.169 × 105 ;
fictitious particles:
2.196 × 105 )
Particle spacing
(initial): 0.2 m
Water 2023, 15, 3130 8 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
Modelled area extent:
17.5 km × 9 km
Unstructured grid
of tetrahedral cells
(diffusion step)
Incompressible Finite element; Historical Number of cells: Flooded areas and Simulation time:
implicit time 1959 Malpasset dam 1.716 × 106 flow velocity fields
Comparison with >8 min
Navier–Stokes 2011
Caboussat et al. [75] N/A splitting scheme break (France) Spatial resolution: at selected times; maximum Run time/ Research
equations coupled physical model data
(advection and Water volume: 5m flood depths and arrival times simulation time:
with VOF 50 million m3 Structured grid 600
diffusion steps) at selected points
of cubic cells
(advection step)
Number of cells: N/A
Spatial resolution:
2m
Modelled area extent:
28.9 km × 5.75 km
Unstructured grid
of tetrahedral cells
(diffusion step)
Finite element; Hypothetical Flooded areas and
Incompressible Number of cells: Simulation time:
implicit time Grande-Dixence 13.876 × 106 flow velocity fields 10 min
Caboussat et al. [75] N/A Navier–Stokes splitting scheme dam break Spatial resolution: at selected times; Inundation Run time/ 2011 Research
equations coupled (Switzerland) dynamics
(advection and 50 m flood depth contour maps at simulation time:
with VOF Water volume: Structured grid N/A
diffusion steps) 400 million m3 selected times
of cubic cells
(advection step)
Number of cells: N/A
Spatial resolution:
10 m
Incompressible
Navier–Stokes
equations coupled Hypothetical
Sayano– Modelled area extent: Flood depth
with grid level set Finite difference/ the reservoir and a valley Simulation time:
finite volume; Shushenskaya hydrographs at 100 s
Vassilevski et al. function (for free stretch downstream of 2012
[76] N/A surface tracking); Chorin–Temam- partial dam-break selected points; time series of Dam-break flow Run time/ Research
Yanenko time the dam simulation time:
(Russia) Structured octree the bottom pressure at the
Herschel–Bulkley splitting scheme N/A
Water volume: staggered grid base of the spillway
rheological relation N/A
for
viscoplastic fluids
Modelled area extent:
the reservoir and a valley
Historical 1963
overtopping of the stretch downstream of the
dam Reconstruction of
Vajont dam (Italy) Number of particles: Water surface elevation the wave generated Simulation time:
Volume of the at selected times; maximum by the Vajont 21 min
Navier–Stokes and rockslide: 3.954 × 106 Run time/
Vacondio et al. Weakly (bottom particles: run-up on the reservoir side; rockslide and of the
[77] *
DualSPHysics continuity
compressible SPH 310 million m3 water surface elevation in the dam-overtopping simulation time: 2013 Open-source
equations Stored water 2.144 × 106 ; residual lake; flow velocity
177
rockslide particles: phenomenon Parallelization
volume:
field; overflow hydrograph Comparison with on GPU
N/A 1.274 × 105 ;
(reservoir water field observations
fluid particles:
level provided)
1.683 × 106 )
Particle size: 5 m
Water 2023, 15, 3130 9 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
Finite volume;
PIMPLE Hypothetical Modelled area extent:
RANS Simulation time:
Zhainakov and algorithm; Andijan dam break 6 km × 4 km Contour maps of the
coupled with VOF; Structured mesh of Flood wave 240 s Open-
Kurbanaliev [78]; OpenFOAM explicit Euler (Uzbekistan) water volume fraction at Run time/ 2013
standard k-ε hexahedral cells propagation source
Jainakov et al. [79] first-order Water volume: selected times simulation time:
turbulence model Number of cells:
time discretization N/A 135
120 × 120 × 80
method
Finite volume;
PIMPLE Hypothetical Modelled area extent:
RANS algorithm; 5 km × 5 km Simulation time:
Zhainakov and coupled with VOF; Papan dam break Contour maps of the
OpenFOAM Structured mesh of Flood wave 260 s 2013 Open-
Kurbanaliev [78]; explicit Euler (Kyrgyzstan) water volume fraction at source
standard k-ε hexahedral cells propagation Run time/
Jainakov et al. [79] first-order time Water volume: selected times
turbulence model Number of cells: simulation time: 69
discretization N/A 50 × 60 × 30
method
Flooding
Hypothetical break Modelled area extent: dynamics
of the embankment the reservoir and a Comparison with
4.5 km long valley stretch Water surface elevation at 2D depth-averaged
of the reservoir of Simulation time:
downstream selected times; transverse
Navier–Stokes and the Kolarjev vrh model predictions 200 s
Weakly Number of particles: water surface profiles at
Džebo et al. [80] Tis Isat continuity pumped-storage and physical model Run time/ 2014 Research
compressible SPH given cross-sections; flow
equations hydropower plant (a) 21.890 × 103 ; experimental data simulation time:
depth hydrographs at selected (a) 30; (b) 981
(Slovenia) (b) 174.884 × 103 gauge points Effects of different
Water volume: Particle size: bottom roughness
3.1 million m3 (a) 5 m; (b) 2.5 m values and spatial
resolutions
Modelled area extent:
Flash flood in 5 km long reach Comparison with
RANS Simulation time:
coupled with VOF; the Toce River 1:100 of the Toce River experimental
Finite volume; physical model 3 min
Marsooli and Wu Smagorinsky eddy (5.5 km × 1.2 km) Flow-depth hydrographs at physical model data
N/A PISO algorithm; Run time/ 2014 Research
[50] (Italy) Unstructured mesh selected points Comparison with
viscosity turbulence CICSAM scheme simulation time:
Controlled of hexahedral cells 2D depth-averaged 4100
model impulsive inflow Number of cells: model predictions
3.1 × 106
Modelled area extent:
5 km long reach
of the Toce River
(50 m × 11 m in the physical
Flash flood in model;
RANS the Toce River 1:100 presence of an idealised urban Validation Simulation time:
coupled with VOF; physical model district) Water depth hydrographs at (comparison with 60 s
Finite volume; 2014
Zhou et al. [81] N/A PISO algorithm (Italy) Structured mesh physical model Run time/ Research
k-ε turbulence selected gauge points simulation time:
model Controlled of prismatic cells experimental data) N/A
impulsive inflow Number of cells:
8.904 × 103
Spatial resolution:
1 m (horizontal)
10−2 m (vertical)
Water 2023, 15, 3130 10 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
Hypothetical
Dongwushi dam VOF spatial
Modelled area extent: distribution at
break upper reach of
(China) selected times; flow velocity Simulation time:
RANS the valley of the Flood wave
coupled with VOF; Water volume: field at selected times; flow ~47 h
Finite volume; Fuyang River propagation
Zhou et al. [81] N/A
k-ε turbulence PISO algorithm 161.5 million m3 discharge hydrographs at Run time/ 2014 Research
Unstructured mesh Dam-break risk simulation time:
model (Hypothetical dam selected cross-sections
of hexahedral cells analysis N/A
break of four other (including the dam site); flood
Number of cells:
dams in the same 79.513 × 103 depth spatial distribution at
Haihe River selected times
basin, China)
Arrival time at
Modelled area extent: selected points; Comparison with
Historical 17.5 km × 10 km flood hydrographs experimental Simulation time:
RANS
coupled with VOF; Finite volume; 1959 Malpasset dam Unstructured mesh at selected points; flooded (field and physical 40 min
Biscarini et al. [82] OpenFOAM PISO algorithm; break (France) Number of cells: area at selected times; model) data Run time/ 2016 Open-source
k-ε turbulence 2.203 × 106 simulation time:
model MULES scheme Water volume: transverse free surface 3D effects at
50 million m3 Spatial resolution: profiles at a river bend sharply curved N/A
N/A cross-sections; velocity fields river bends
in selected areas
Modelled area extent:
17 km × 9 km
Unstructured
horizontal mesh
Navier–Stokes and Historical of triangular elements Flood depth contour maps at Simulation time:
TELEMAC continuity Finite element; 1959 Malpasset dam selected times; 4000 s
equations Number of cells: Flood wave
Modelling TELEMAC-3D three-fractional-step break (France) (a) 26 × 103 ; flood depth Run time/ 2016 Freeware
(Boussinesq propagation simulation time:
System [83] algorithm Water volume: hydrographs at
50 million m3 (b) 104 × 103 N/A
approximation) Vertical mesh: selected locations
2 or 6 layers regularly spaced
in the vertical
direction
Continuity
equations of the
fluid and solid Erosional Modelled area extent: 3D distribution of the
incompressible dam-break 24.627 km × 9.855 km particles and velocity fields at
phases + volume demonstrative Mobile bottom downstream selected times; maps of Simulation time:
Weakly- of the dam (granular material
balance equation; compressible SPH; ICOLD maximum values of mixture Dynamics of the 25 min
Amicarelli et al. [84] SPHERA momentum benchmark of fixed characteristics) depth and specific flow rate; Run time/ 2017 Open-source
Leapfrog time Water volume: Number of particles: phenomenon simulation time:
equations of the water and bed-load flow rate
integration scheme N/A N/A
fluid and solid 6.8 × 105 and mixture depth
phases; (reservoir water Particle spacing: hydrographs at selected
momentum level provided) 4m cross-sections
equation for the
mixture
Water 2023, 15, 3130 11 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
Validation
Modelled area extent: (comparison with
5 km long reach physical model
of the Toce River experimental data)
Flash flood in (50 m × 11 m in the physical
the Toce River 1:100 Computational time; Dam-break flooding Simulation time:
RANS model; two idealised urban
coupled with VOF; physical model of an urban area 60 s
Finite volume; district configurations) computational error; water 2017
Wang et al. [85] N/A Comparison of Run time/ Research
k-ε turbulence PISO algorithm (Italy) Unstructured mesh depth hydrographs at selected simulation time:
model computational
Controlled of polyhedral cells gauge points ~20
impulsive inflow performance of
Number of cells:
~2 × 105 different mesh
Spatial resolution: types (polyhedral,
0.1 m tetrahedral,
hexahedral)
VOF spatial
distribution at
Hypothetical dam Modelled area extent: selected times; flood depth
RANS 40.12 km2 area and flow Simulation time:
break of an urban
coupled with VOF; Unstructured mesh velocity hydrographs at Dam-break flooding 5h
Finite volume; reservoir (SZ City, 2017
Wang et al. [85] N/A of polyhedral cells Run time/ Research
k-ε turbulence PISO algorithm China) selected sites in the urban of an urban area simulation time:
model Water volume: Number of cells: area; velocity and vorticity N/A
94 million m3 4.229 × 106 fields; maximum flood depth
and flow velocity contour
maps
Modelled area extent:
N/A
Historical 2015 (the pond and Flow fields (velocity Simulation time:
Fundão tailings the area around) Tailings flow 30 min
magnitude) and flooded areas
Navier–Stokes and Weakly dam break Number of particles: dynamics Run time/
Wang et al. [86] DualSPHysics continuity (Brazil) at selected times; simulation time: 2018 Open-source
equations compressible SPH Released tailings 2.988 × 106 (fluid) flow depth, velocity, and Comparison with N/A
18.132 × 106 impact pressure time series at field data Parallelization
volume:
32 million m3 (boundaries) a selected location on GPU
Particle spacing:
3m
Modelled area extent:
Hypothetical dam N/A Simulation time:
break of an (the pond and Flow fields (velocity
magnitude) and flooded areas 10 min
Navier–Stokes and operating overhead the area around) Run time/
Weakly tailings pond Number of particles: at selected times; Tailings flow
Wang et al. [86] DualSPHysics continuity simulation time: 2018 Open-source
compressible SPH flow depth, velocity, and dynamics
equations (China) 4.463 × 106 (fluid) impact pressure time series at
N/A
Pond capacity: 3.9 × 106 (boundaries) Parallelization
a selected location on GPU
33 million m3 Particle spacing:
3m
Water 2023, 15, 3130 12 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
Modelled area extent:
100 m × 100 m
Unstructured
Hypothetical horizontal mesh
Navier–Stokes and of triangular elements; Simulation time:
continuity dike-break flooding
Finite element; Flood wave 7 min
equations θ time-stepping on a realistic spatial resolution: 5 m Velocity fields and flooded
Zhang et al. [87] N/A propagation Run time/ 2018 Research
(Boussinesq topography Vertical mesh: areas at selected times simulation time:
method 3D effects
(fixed inflow 1 layer N/A
approximation)
velocity) Unstructured mesh
of tetrahedral cells
Number of cells:
3.114 × 103
Modelled area extent:
5.495 km × 2.5 km
Unstructured
horizontal mesh
Hypothetical of triangular elements;
Navier–Stokes and dike-break flooding spatial resolutions: Flooded areas, flood depth Simulation time:
continuity Finite element; (a) 30–50 m; 3D effects 6h
in a realistic contour maps, and velocity
Zhang et al. [87] N/A equations θ time-stepping urban area (b) 60–100 m; Sensitivity to the Run time/ 2018
fields at selected times; flood simulation time: Research
(Boussinesq (flow velocity (c) 120–200 m
method depth hydrographs at selected mesh resolution (a) 17.5; (b) 3.5;
Vertical mesh:
approximation) through the breach: 1 layer points (c) 0.3
0.1 m/s) Unstructured mesh
of tetrahedral cells
Number of cells:
(a) 9376; (b) 3024;
(c) 816
Debris flow
dynamics
Flow fields (velocity Fluid–structure
magnitude) and flooded areas interactions
Modelled area extent: Comparison with
Historical 1985 the pond and a at selected times;
Navier–Stokes and Stava tailings dam average velocity profile of the other numerical
continuity 4.2 km long stretch results Simulation time:
break of the valley debris flow front; velocity Effect of the N/A
LS-DYNA equations; material SPH (Italy) 2019
Chen et al. [88] field near the check dam; presence of Run time/ Commercial
with Number of particles: simulation time:
Released tailings arrival time at an observation hypothetical
fluid-elastoplastic 11.119 × 103 (fluid) point; final deposition zones; N/A
properties volume: check dams (rigid
185 × 103 m3 Particle spacing: impact force hydrographs
2.5 m indestructible dams
(considering a single or or concrete
multiple check dams) destructible dams)
placed at
different positions

Finite volume; Hypothetical


Modelled area extent:
RANS PISO algorithm; dam-break flow in Maps of the water volume Simulation time:
~8 km × 3 km
coupled with VOF; the Willow Creek 400 s
Kurbanaliev et al. OpenFOAM explicit Euler Mesh of fraction at selected times; Flood wave Run time/ 2019 Open-source
[89] standard k-ε Mountain area hexahedral cells flood depth hydrographs at propagation
first-order (California) simulation time:
turbulence model time discretization Number of cells: selected points
Water volume: ~45
method 0.45 × 106
N/A
Water 2023, 15, 3130 13 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
RANS
coupled with VOF; Modelled area extent:
three 17 × 103 m2 ,
incompressible 1.317 km long river reach
phases for the Hypothetical downstream of the dam
Homogeneous mud layer of Flood depth hydrographs at Flood wave Simulation time:
simulation of mixed Mynzhylky (with mud)
fixed thickness downstream selected points; water 60 s
Issakhov and water–mud flow: Finite volume; erosional dam break propagation 2020
N/A of the dam surfaces and inundated areas Run time/ N/A
Zhandaulet [90] two Newtonian PISO algorithm (Kazakhstan) Structured mesh at selected times (for different Effect of the initial simulation time:
fluids (air and Water volume: mud layer thickness N/A
of tetrahedral cells mud layer thicknesses)
water) and a 50 × 103 m3
non-Newtonian Number of cells:
liquid; 2.433 × 106
Spatial resolution:
realizable k-ω 0.5 m
turbulence model
Modelled area extent:
18 km long river reach Flood wave
downstream of the dam and Flooded areas at propagation
floodplains different times; 3D effects
Hypothetical free surface profile along the Comparison with
Lake: river at Simulation time:
Coralville unstructured grid 2D depth-averaged 5h
RANS dam-break peak flood extent; discharge
coupled with VOF; Finite volume; with polyhedral cells; hydrographs at selected river model predictions Run time/
Munoz and STAR-CCM+ (USA) 2020
SIMPLE spatial resolution: sections; Recalibration of the simulation time: Commercial
Constantinescu [37] realizable k-ε Water volume:
algorithm 100 m unit discharge transverse 2D model 144
turbulence model N/A Parallelization
(reservoir water River and floodplains: profiles in selected parameter to
unstructured grid improve the using MPI
level provided) cross-sections at peak flood
with prismatic cells; extent; details of the agreement between
multi-resolution velocity field 2D and 3D model
Number of cells: results
18 × 106
Modelled area extent:
18 km long river reach
downstream of the dam and
floodplains
Hypothetical Lake: Flooded areas at Flood wave Simulation time:
Saylorville dam unstructured grid different times;
RANS propagation 3.75 h
Finite volume; break with polyhedral cells; free surface profile along the Run time/
Munoz and coupled with VOF; (USA) 3D effects
STAR-CCM+ SIMPLE spatial resolution: river at the end of the Comparison with simulation time: 2020 Commercial
Constantinescu [37] realizable k-ε Water volume:
algorithm N/A simulation; discharge 230
turbulence model N/A 2D depth-averaged Parallelization
River and floodplains: hydrographs at selected river
(reservoir water model predictions using MPI
unstructured grid sections
level provided)
with prismatic cells;
multi-resolution
Number of cells:
40 × 106
Modelled area extent:
Navier–Stokes and the pond and Simulation time:
Hypothetical
continuity the area around 10 min
equations; Yujiaquan tailings (~2 km × 2 km) Flow fields (velocity Run time/
Weakly dam break Number of particles: Tailings flow
Wang et al. [91] DualSPHysics generalised magnitude) and flooded areas simulation time: 2020 Open-source
compressible SPH (China) dynamics
Herschel–Bulkley– Pond capacity: 3.495 × 106 (fluid) at selected times N/A
Papanastasiou 0.936 × 106 Parallelization
rheological model 52.55 million m3 on GPU
(boundaries)
Particle spacing: 2 m
Water 2023, 15, 3130 14 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
Historical 2019 Modelled area extent: Simulation time:
Feijão N/A 2500 s
RANS (Brumadinho) (suitable area around the Flow velocity
coupled with VOF; Run time/
tailings dam break reservoir) Tailings flow simulation time:
magnitude contour maps at
standard k-ε (Brazil) Unstructured mesh N/A
Finite volume; selected times; wave front dynamics
Yu et al. [92] OpenFOAM turbulence model; Pond capacity: of hexahedral cells Parallelization 2020 Open-source
Bingham– PISO algorithm Number of cells: motion; free surface average Comparison with
12.7 million m3 using MPI (analysis
Papanastasiou 3.242 × 106 velocity hydrograph; flooded field data
Released tailings Spatial resolution: area of the speed-up of
rheological model
volume: 10 m (horizontal) different numbers
11.7 million m3 3 m (vertical) of processors)

Simulation time:
Modelled area extent: 800 s
RANS N/A Run time/
Hypothetical (suitable area around the Flow velocity
coupled with VOF; simulation time:
A’xi gold tailings reservoir) magnitude contour maps at N/A
standard k-ε Tailings flow
OpenFOAM turbulence model; Finite volume; dam break Unstructured mesh selected times; wave front Parallelization 2020 Open-source
Yu et al. [92] (China)
Bingham– PISO algorithm of hexahedral cells; motion; free surface average dynamics using MPI
Papanastasiou Pond capacity: number of cells: velocity hydrograph; flooded (analysis of the
3.6 million m3 6.657 × 106 area
rheological model Spatial resolution: speed-up of
3m different numbers
of processors)
Historical
landslide dam Landslide and
break following landslide
consequent to the Modelled area extent: Flood depth contour maps at Simulation time:
RANS 2000 Yigong ~33 km long stretch selected times; flow depth dam-break coupled
coupled with VOF; 3 h 20 min
Zhuang et al. [93] FLOW-3D Finite volume landslide (China) of the Yigong River valley and velocity hydrographs at 3D simulations Run time/ 2020 Commercial
RNG k-ε Water volume: Comparison with simulation time:
turbulence model Mesh details: selected points; flow
N/A N/A field data N/A
discharge at selected sections Flood wave
(water depth
of 60 m at the propagation
barrier lake)
Urban flood
Flooded areas; features
Hypothetical Modelled area extent: velocity fields Comparison with Simulation time:
Alpe Gera dam 7.9 km × 9.9 km
Euler and Weakly- Number of particles: at selected times; experimental 50 min
SPHERA break maximum flood 2021 Free and
Amicarelli et al. [94] continuity laboratory data Run time/
compressible SPH (Italy) N/A depth contour map; discharge open-source
equations simulation time:
Water volume: Particle spacing: Adoption of a
and flood depth hydrographs N/A
68.1 million m3 N/A flooding damage
at selected sections
model
Flow depth
Modelled area extent: hydrographs at
Hypothetical N/A selected sites;
RANS Attabad Lake (stretch of the flow discharge Simulation time:
coupled with VOF; landslide dam downstream valley) ~1 h 19 min
Karam et al. [95] FLOW-3D Finite volume hydrographs at selected Flood wave Run time/ 2021 Commercial
RNG k-ε turbulence break Multiple mesh blocks of propagation
(Pakistan) cross-sections; flood simulation time:
model hexahedral cells inundation maps and velocity N/A
Water volume:
305 million m3 Number of cells: fields at selected times; flood
N/A arrival times at
selected locations
Water 2023, 15, 3130 15 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
Flash flood in
the Toce River 1:100
physical model
Lattice Boltzmann
(Italy)
equation with the Simulation time:
Controlled Modelled area extent: Video animations of the
Bhatnagar–Gross– impulsive inflow numerical results; flood depth N/A
Lattice Boltzmann 5 km long reach Flood wave
Miliani et al. [96] N/A Krook (BGK) Run time/ 2021 Research
algorithm Two case studies of the Toce River hydrographs at propagation
collisional operator; simulation time:
considered: the (5.5 km × 1.2 km) selected gauge points
interface tracking presence of actual N/A
method buildings and an
idealised array of
buildings
Modelled area extent:
Flash flood in 5 km long reach
the Toce River 1:100 of the Toce River
physical model (5.5 km × 1.2 km) Simulation time:
RANS (Italy) Unstructured mesh Non-hydrostatic 1 min
coupled with a Controlled of prismatic cells effects Run time/
Coupled
free-surface impulsive inflow with triangular basis Comparison with simulation time:
equation; finite volume– 29.4
finite difference; (high-inflow and in a vertical Flow depth hydrographs at experimental
Ai et al. [97] N/A non-hydrostatic and (high-inflow 2022 Research
explicit low-inflow boundary-fitted selected points physical model data
hydrostatic hydrographs) hydrograph)
projection coordinate system Comparison with
versions; Array of aligned Run time/
method Number of cells: hydrostatic 3D simulation time: 24
standard k-ε
turbulence model buildings 1.041 × 105 model predictions (low-inflow
simulating a (2.082 × 104 triangular hydrograph)
simplified urban elements on the
district bottom and 5 layers along the
vertical)
RANS
coupled with VOF;
three Modelled area extent:
incompressible Hypothetical N/A
phases for the erosional (a stretch of the river) Water surfaces and
inundated areas at different Flood wave Simulation time:
simulation of mixed dam-break flow Homogeneous mud layer of
times; flood depth (with mud) 34.5 s
water–mud flow: Finite volume; along the fixed thickness downstream
Issakhov et al. [98] ANSYS Fluent propagation Run time/ 2022 Commercial
two Newtonian PISO algorithm Kargalinka River of the dam hydrographs at selected
Effect of the initial simulation time:
fluids (air and (Kazakhstan) Structured mesh points (for different mud layer N/A
mud layer thickness
water) and a Water volume: of uniform cells thicknesses)
non-Newtonian 333.5 × 103 m3 Number of cells:
liquid; 0.985 × 106
realizable k-ω
turbulence model
Hypothetical Modelled area extent: Flow fields (velocity
Dagangding N/A magnitude) and flooded areas
RANS
coupled with VOF; tailings dam break (selected area downstream of at selected times; wave front Simulation time:
(China) the dam) advancement and 2000 s
standard k-ε Finite volume; celerity in time; final Tailings flow
Yang et al. [99] ANSYS CFX Pond capacity: Unstructured mesh with Run time/ 2022 Commercial
turbulence model; PISO algorithm tetrahedral and pentahedral dynamics
Bingham 3.59 million m3 deposition area and depth simulation time:
(Theoretical inflow cells distribution; longitudinal and N/A
rheological model Number of cells:
discharge transverse profiles
at the dam site) 0.543 × 106 of the final deposit
Water 2023, 15, 3130 16 of 26

Table 2. Cont.

(4) (6) (8) (9)


(1) (2) (3) (5) Computational (7) Computational (10) (11)
Numerical Focus of
Reference Model Name Model Type 1 Case Study Domain and Output Data Year Status
Method 2 the Study Efficiency 3
Elements
Propagation of
Historical the tailings slurry
2017 Tonglüshan Modelled area extent: Comparison with
~2 km × 1.5 km field data
tailings dam break area around the Flow depth maps at different Simulation time:
Hydrodynamic concerning the final
equations; (China) tailings pond times; final deposition area 300 s
Zhuang et al. [100] DAN3D SPH Pond capacity: deposition Run time/ 2022 Freeware
rheological models Number of particles: and slurry depth distribution;
distribution of the simulation time:
(Bingham model) 15.78 million m3 ; 4 × 103 maximum velocity magnitude
tailings slurry
moved slurry map N/A
Particle spacing: Sensitivity analysis
volume: N/A
0.5 million m3 on the solid
concentration of the
tailings slurry

Notes: 1 RANS = Reynolds-Averaged Navier–Stokes equations; RNG = Re-normalization Group; VOF = Volume of Fluid. 2 CICSAM = Compressive Interface Capturing Scheme for
Arbitrary Meshes; MULES = Multidimensional Universal Limiter with Explicit Solution; PIMPLE = combination of PISO and SIMPLE; PISO = Pressure Implicit with Split Operators;
SIMPLE = Semi-Implicit Method for Pressure-Linked Equations; SPH = Smoothed-Particle Hydrodynamics. 3 GPU = Graphics Processing Unit; MPI = Message Passing Interface.
N/A = not available; * dam spillway or overtopping flow.
Water 2023, 15, x FOR PEER REVIEW 19 of 29

Water 2023, 15, 3130 17 of 26


3. Results and Discussion
A total of 34 documents relating to the period from 2006 to the present were retrieved
3. Results and Discussion
from the literature review. However, 39 entries appear in Table 2, as nine of the references
A totalpresent
examined of 34 documents
two different relating
case to the period
studies, hencefrom
being2006 to the present
repeated in two were retrieved
different table
from the literature review. However, 39 entries appear in Table 2, as
rows, and four references are mentioned together with other works that contain identical nine of the references
examined present two different case studies, hence being repeated in two different table
case studies.
rows, and four references are mentioned together with other works that contain identical
Figure 3 summarises the main statistical information derived from the analysis of the
case studies.
papers reviewed. About a third of these (12) were published in the last three years (2020–
Figure 3 summarises the main statistical information derived from the analysis of
2022), confirming a recent growing interest in 3D modelling of real-field dam-break flows
the papers reviewed. About a third of these (12) were published in the last three years
(Figure 3a). In most case studies examined (43.6%), in-house research codes were adopted
(2020–2022), confirming a recent growing interest in 3D modelling of real-field dam-break
for the numerical analyses; commercial CFD software packages and open-source or
flows (Figure 3a). In most case studies examined (43.6%), in-house research codes were
freeware codes were used in 17.9% and 35.9% of cases, respectively (Figure 3b). The type
adopted for the numerical analyses; commercial CFD software packages and open-source
of code was unspecified in only one case. Mesh-based methods appear in Table 2 more
or freeware codes were used in 17.9% and 35.9% of cases, respectively (Figure 3b). The
times (25) than particle-based ones (14) (Figure 3c). Among the former methods, the VOF-
type of code was unspecified in only one case. Mesh-based methods appear in Table 2
based finite volume ones are the most commonly used, whereas, among the latter, the
more times (25) than particle-based ones (14) (Figure 3c). Among the former methods, the
SPH methods
VOF-based arevolume
finite prevalent.
onesThis state
are the mostof the art was used,
commonly also observed
whereas, by otherthe
among researchers
latter, the
(e.g., [101]) who, in addition, showed how numerical results obtained
SPH methods are prevalent. This state of the art was also observed by other researchers from these two
methods
(e.g., arewho,
[101]) in good agreement
in addition, [102]. The
showed howLattice Boltzmann
numerical resultsmethod
obtained appears in Table
from these two2
in only one case. This modern numerical technique still has limited
methods are in good agreement [102]. The Lattice Boltzmann method appears in Table 2 in application to dam-
breakone
only problems
case. This [103,104].
modern Finite volume
numerical techniques
technique arelimited
still has the most used in to
application thedam-break
reviewed
problems [103,104]. Finite volume techniques are the most used in the reviewed papers3d).
papers for discretizing the governing equations in mesh-based models (Figure for
Indeed, the the
discretizing finite volume equations
governing method isinparticularly
mesh-basedsuitable for solving
models (Figure 3d). conservation
Indeed, the finite law
equations
volume because
method it exploits the
is particularly integral
suitable forformulation of the equations
solving conservation to capture
law equations their
because
weak solutions [105]. Finite element methods appear fewer times in
it exploits the integral formulation of the equations to capture their weak solutions [105]. Table 2, since this
numerical technique still constitutes a developing research area in
Finite element methods appear fewer times in Table 2, since this numerical technique still CFD [106] and free
surface flow
constitutes modelling [107],
a developing despite
research area initsCFD
ability toand
[106] treatfree
complex
surfacegeometries and [107],
flow modelling reach
high-order accuracy.
despite its ability to treat complex geometries and reach high-order accuracy.

Figure 3. Statistical information from the database of the reviewed papers on 3D modelling of large-
Figure 3. Statistical information from the database of the reviewed papers on 3D modelling of
scale real-world dam-break floods: (a) year of publication; (b) status of the software; (c) model type;
large-scale real-world
(d) numerical scheme.dam-break floods: (a) year of publication; (b) status of the software; (c) model
type; (d) numerical scheme.
The catastrophic scenario of the sudden and total dam collapse is mostly considered
The catastrophic scenario of the sudden and total dam collapse is mostly considered
in the case studies reviewed, thus neglecting the breach development dynamics.
in the case studies reviewed, thus neglecting the breach development dynamics.
Three-dimensional models employed in dam-break modelling are also commonly
used to simulate floods in rivers (e.g., [45,108]) or inundations in floodplains [19] or in
urban areas (e.g., [36,109]).
Water 2023, 15, 3130 18 of 26

3.1. Improvements in Simulation Accuracy


Three-dimensional hydrodynamic modelling improves the mathematical description
of dam-break flows compared to routinely adopted 2D depth-averaged modelling, because
3D models overcome the intrinsic limitations of 2D ones. Therefore, 3D modelling is of
wide applicability and general interest in dam-break problems. Indeed, 3D models calculate
the pressure field and include the vertical fluid acceleration, thereby inherently taking into
account the effects of flow curvature. Moreover, 3D models involve the vertical velocity
component and describe the vertical variation of flow velocity. Conversely, 2D depth-
averaged models introduce hypotheses on the vertical pressure distribution (assumed
hydrostatic in SWE models) or the shape of the vertical velocity profile in non-hydrostatic
flow models [23]. Finally, 3D models can include vertical turbulence and spiral flows [60]
and can correctly simulate the impact of flows against obstacles or structures [109,110].
These enhanced modelling capabilities of 3D models can ensure more accurate flow pre-
dictions, especially around structures, at topographic singularities where sudden changes
in the bottom surface occur, and in urban flooding simulations. Reproducing small-scale
eddies, vertical turbulence, and residual circulation can be crucial for accurately predict-
ing both near-field and far-field features of the flooding dynamics and the flow variables
involved in flood hazard assessments.
The capabilities of 3D models were verified and validated in most of the articles reviewed,
but the analysis of 3D flow effects is the focus of only a few of them (e.g., [70,71,82,87]). For
example, Biscarini et al. [82] discussed the 3D effects induced on a dam-break flow by a
sharp river bend. Even fewer are the articles that, based on a real-field case study, compare
the results obtained through a 3D model with those obtained through a standard 2D depth-
averaged shallow-water model, analysing the differences [37,50,80]. For example, Munoz and
Constantinescu [37], studying the flood inundations induced by the hypothetical failure of
two flood-protection dams in the United States, found that their 2D depth-averaged model
underpredicted the wave propagation speed and the inundation extent compared with a 3D
model. Most frequently, in the literature, the comparison of 3D and 2D model performance in
predicting dam-break flow was made on the basis of schematic test cases characterised by a
simple geometry (e.g., [27,111]).
In 3D models based on the RANS equations, the effect of turbulence is introduced
through a closure turbulence model. The classic k-ε model is the most adopted in the
references reviewed (entries [37,82,85,95] in Table 2). However, no systematic sensitivity
analyses on the turbulence model type were performed for real-world dam-break case
studies. The prediction capabilities of different turbulence models were typically compared
again on the basis of dam-break test cases characterised by a simple and schematic geometry
(e.g., [110,112–116]), exploiting laboratory experimental data.
Some references analysed in this review consider mudflows resulting from a tailings
dam failure. In these studies, the tailings slurry is assumed as a homogeneous non-
Newtonian viscoplastic fluid, and a suitable constitutive equation is added to the set of
governing equations to characterise its rheological properties [88,91,92,99,100]. In some
other studies among those reviewed, non-Newtonian rheological models are used to
describe the behaviour of moving sediment layers in the path of the dam-break wave [90,98],
thereby simulating the geomorphic effects produced by a dam-break flood on an erodible
bed. An erosional dam break (with bed-load transport) on real-world topography was
modelled by Amicarelli et al. [84] through an SPH model applied to a mixture of water and
a non-cohesive granular material.
Recently, topographic data accuracy and spatial resolution have been significantly
improved thanks to highly accurate terrain surveying techniques, such as scanning airborne
laser altimetry (LIDAR), which can reach a horizontal resolution even below 1 m and a
vertical accuracy of 10−1 m [19], and unmanned aerial vehicle (UAV) photogrammetry,
which is very effective in terms of timeliness, repeatability, and high resolution [91]. Hence,
highly accurate and high-resolution topographic digital models are widespread nowa-
days, even in urban areas [85], and allow for very accurate dam-break flow modelling.
Water 2023, 15, 3130 19 of 26

On the other hand, despite the implicit greater descriptive capacity (and complexity) of
3D models, accuracy improvement in 3D numerical predictions compared with simpler,
low-dimensional models may be illusory if topographic and input data are limited and
inaccurate and reliable real-field validation data are scarce [19,117].

3.2. Model Validation and Calibration


Dam-break models are usually validated by comparing numerical results with experi-
mental data [13]. To this end, real-field data of historical dam-break events or experimental
data obtained from small-scale physical models with an irregular bottom can be very
useful [14], especially if the numerical model will be used in practical engineering applica-
tions for dam-break inundation mapping and flood hazard assessment. Validation against
real-field data can also help discriminate among potential competing models able to return
plausible and realistic results.
Based on the articles reviewed, the most used real-field test case for validating 3D
dam-break numerical models is the historical event of the 1959 Malpasset dam break
(entries [75,82,83] in Table 2). An extensive database is available for this test case, including
field data collected immediately after the event and data measured during a laboratory
investigation in a reduced-scale physical model [118,119]. However, other historical dam-
break events are well documented [14]; for some of these, the test case and the experimental
data (from field surveys or consultation of historical documents) are available in digital
format [120,121].
Another test case frequently used for validating 3D dam-break numerical models is
the Toce River test case (entries [50,81,85,96] in Table 2), which concerns a dam-break-like
flash flood induced by imposing an inflow discharge hydrograph in a 1:100 physical model
of a 5 km long stretch of the Toce River valley in northern Italy [122]. An idealised urban
district was inserted in this physical model to simulate the dam-break flooding of an urban
area. Other experimental data from reduced-scale physical models are available in the
literature [123,124] and could be considered for validating 3D dam-break models.
In complex 3D models, a large number of model parameters requiring calibration
appear, mainly concerning numerical schemes and turbulence models. Accordingly, an
optimization process involving many parameters should be performed, which might
be challenging and computationally expensive due to the long runtime of each model
execution. Therefore, calibration parameters are usually set based on expert judgement,
according to the suggestions of software users’ guides or following the choices made in
similar numerical studies.
Uncertainty analysis is widely recognised as desirable, if not indispensable, in environ-
mental system modelling to associate uncertainty estimates with model predictions [15]. In
real-world applications of dam-break inundation modelling, the main uncertainty sources
are topographic input data, model parameters (including the mesh resolution), and initial
conditions defining the dam-break scenario (i.e., breach parameters and reservoir filling
conditions). Modern remote sensing acquisition technologies (such as LIDAR) have drasti-
cally reduced the uncertainty in terrain description, and the availability of high-resolution
topographic data makes it possible to include in the computational domain obstacles, struc-
tures, and topographic details that may significantly influence the inundation process [60].
The typical way to individually quantify the effects of model or scenario parameters on
numerical predictions is to perform a sensitivity analysis based on a sampling approach
(e.g., [125,126]). Monte Carlo methods are usually adopted to this end, considering a sample
of the appropriate size to ensure convergence. However, exhaustive sensitivity analyses
based on large sets of parameter values (and, consequently, many model runs) may be
prohibitive using 3D hydrodynamic models. To overcome this limitation, Rizzo et al. [10]
(who, however, used a 2D depth-averaged shallow-water model) proposed a probabilistic
method based on a limited set of dam-break scenarios, each of which had a (conditional)
probability associated with it. As regards the spatial resolution, Zhang et al. [87] performed
a sensitivity analysis on the mesh size for a realistic dike-break flooding case, concluding
Water 2023, 15, 3130 20 of 26

that the spatial resolution significantly impacts the accuracy of model results when the
terrain is irregular, as in real-field applications.

3.3. Improvements in Computational Efficiency


The main reason for the limited use of 3D models in large-scale dam-break flood
simulation over real-world topography is that they are time-consuming due to their high
computational cost, which depends on various factors, such as the complexity of the nu-
merical method adopted to solve the governing equations, the extent of the computational
domain, the simulation (physical) time (usually set long enough to reconstruct the salient
features of the flooding process), and the mesh type and resolution (in mesh-based models)
or the particle size (in particle-based models). This problem—which is exacerbated in
real-world applications—does not concern, as a rule, dam-break tests characterised by
a schematic geometry (i.e., verification against analytical solutions or validation against
laboratory data), because such simple tests do not require, in general, significant computa-
tional resources.
A viable option to overcome the limitation related to computational efficiency without
resorting to approximate models [127] is to accelerate 3D model calculations via high-
performance computing and GPU technology to reduce model running times [19]. In
particular, parallel computing is a valid and widely used method to improve computational
efficiency, possibly exploiting GPU computing power and processing capabilities.
Reducing the computational time of model executions also allows larger domains
to be considered with high spatial resolution, extending the simulation of the flooding
dynamics for a longer simulation time. Moreover, a large set of dam-break scenarios can be
used in sensitivity analyses for assessing model uncertainty.
Table 2 shows that parallelization techniques based on Message Passing Interface
(MPI; [39,72]) and the GPU implementation of the simulation model [77,86,91] were used
in 3D modelling of real-field dam-break flows. The ratio of model run time to simulation
time is highly variable, ranging from 102 to 104 , depending on the number of processors
and the total number of computational cells or particles. Yu et al. [92] analysed the gain in
parallel speed-up as a function of the number of processors.

3.4. Improvements in Result Visualization


Geographic information systems (GISs; [72,73]) and 3D virtual geographic environ-
ment (VGE) systems have recently become attractive tools for the 3D dynamic visualization
of model results, improving the communication of the dam-break flooding dynamics and
the potential consequences (e.g., [128]). The availability of 3D numerical results facilitates
the development of 3D virtual reality environments and visualizations, which, currently, are
usually based on 2D hydraulic model data and results extrapolated into 3D (e.g., [129,130]).

4. Conclusions
Three-dimensional numerical modelling of dam-break flow has developed consid-
erably in the last decade thanks to the significant increase in the available computing
resources, which has made 3D modelling a viable alternative to routinely used 2D depth-
averaged modelling in large-scale real-field applications, despite its higher computational
cost. Even if 3D numerical models have only recently become a real and feasible option in
large-scale dam-break modelling on real-world topography, we found in the literature a
noticeable number of contributions concerning this fluid dynamic application, including
standard dam-break water flows as well as geomorphic and tailings dam-break flows.
Mesh-based models are mostly used for solving the governing hydrodynamic equations,
which are the Navier–Stokes equations or the RANS equations coupled with a free-surface
tracking technique (e.g., the VOF method) and a closure turbulence model. However, in
recent years, particle-based models based on the SPH technique have become widespread
in computational fluid dynamics and in the simulation of dam-break flows. Indeed, this
method benefits from both being mesh-free and not requiring computationally expensive
Water 2023, 15, 3130 21 of 26

free-surface tracking techniques. The application of the Lattice Boltzmann method to


large-scale 3D dam-break flood modelling is a relatively new research field that deserves
further exploration [96].
This paper systematically reviews the state-of-the-art 3D numerical modelling of
large-scale dam-break floods on irregular real-world topography. The literature survey
is mainly based on journal and conference papers published until July 2023, excluding
the grey literature. We aimed to conduct an exhaustive and meticulous review based on
comprehensive search parameters and inclusive keywords. Nevertheless, we may have
missed studies published in journals of local diffusion or not specifically focused on 3D
numerical modelling.
The references reviewed are organised into a table (Table 2) reporting extensive infor-
mation on numerical models, case studies analysed, research focus, numerical results, and
computational efficiency. Recent developments and key improvements in modelling and
computational aspects, such as model accuracy and efficiency, are discussed. Regarding
computational efficiency, code running on Graphics Processing Units and massive paral-
lelization ensure a significant computational time reduction. This general improvement in
computational efficiency allows for high spatial resolutions, even in large-scale applications.
However, model calibration remains challenging due to the large number of parameters
involved in 3D models, especially when coupled with a turbulence model. The 3D visu-
alization of the numerical results can improve the quality of the communication of the
dam-break flood hazard to managers and stakeholders, thus contributing to the mitigation
of dam-break flood consequences.
Compared to the less computationally demanding 2D depth-averaged models, the 3D
ones allow for a more detailed and accurate prediction of the flooding dynamics, inherently
including the fluid vertical acceleration and 3D effects due to the flow curvature induced by
the irregular topography. Freeware and commercial CFD software (relatively user-friendly)
are available nowadays for dam-break flow analysis, even concerning large-scale problems
on real-world topography.
Future research may concern coupled 2D–3D models as a valid compromise between
simulation accuracy and computational efficiency. This modelling option is currently
already introduced in some CFD software (e.g., [131]) in order to simulate large-scale flows
in which the shallow water assumptions are approximately valid with a 2D depth-averaged
model, and near-field flows and localised flow features (near structures, obstacles, or
significant topographic irregularities) with a 3D model. For instance, hybrid 2D–3D models
could be used to simulate dam-break flooding generated by a partial dam failure, with the
breach involving only the upper part of the dam. In this case, the weir-type outflow could
be modelled by the 3D model and the wave propagation in the downstream area by the 2D
depth-averaged model. The portions of the computational domain in which to apply the
two different models can be previously identified and efficiently linked.
This review can guide researchers, modellers, and practitioners to compare existing
3D dam-break numerical models, choose the most suitable model for the application of
interest, and select state-of-the-art numerical approaches. This review can also support
modellers and researchers, providing a basis for future research in 3D simulation models
and computational techniques for dam-break flow modelling, with special attention to
large-scale real-field applications.

Author Contributions: A.M. and M.T. contributed equally to the conceptualization and implementa-
tion of the research, the revision of the existing literature, the analysis of results, and the writing of
the manuscript. All authors have read and agreed to the published version of the manuscript.
Funding: A.M. was supported by the Italian Ministry of University and Research through the PRIN
2017 Project RELAID (REnaissance of LArge Italian Dams), project number 2017T4JC5K.
Data Availability Statement: No new data were created or analysed in this study. Data sharing is
not applicable to this article.
Conflicts of Interest: The authors declare no conflict of interest.
Water 2023, 15, 3130 22 of 26

References
1. Costa, J.E. Floods from Dam Failures; Open-File Report 85-560; US Geological Survey: Denver, CO, USA, 1985. Available online:
https://pubs.usgs.gov/of/1985/0560/report.pdf (accessed on 21 July 2023).
2. Charles, J.A.; Tedd, P.; Warren, A. Lessons from Historical Dam Incidents; Project SC080046/R1; Environment Agency: Bristol, UK,
2011. Available online: https://assets.publishing.service.gov.uk/media/603369e7e90e07660cc43890/_Lessons_from_Historical_
Dam_Incidents_Technical_Report.pdf (accessed on 21 July 2023).
3. Zhang, L.; Peng, M.; Chang, D.; Xu, Y. Dam Failure Mechanisms and Risk Assessment; Wiley: Singapore, 2016.
4. Graham, W.J. A Procedure for Estimating Loss of Life Caused by Dam Failure; DSO-99-06; US Department of Interior, Bureau of
Reclamation, Dam Safety Office: Denver, CO, USA, 1999. Available online: https://www.usbr.gov/ssle/damsafety/TechDev/
DSOTechDev/DSO-99-06.pdf (accessed on 21 July 2023).
5. ASDSO (Association of State Dam Safety Officials). Lesson Learned from Dam Incidents and Failures—Case Studies. Available
online: https://damfailures.org/case-study/ (accessed on 21 July 2023).
6. International Water Power and Dam Construction. Learning Historical Dam Safety Lessons. Available online: https://www.
waterpowermagazine.com/features/featurelearning-historical-dam-safety-lessons-4958949/ (accessed on 21 July 2023).
7. FEMA. Federal Guidelines for Inundation Mapping of Flood Risks Associated with Dam Incidents and Failures; FEMA P-946; Federal
Emergency Management Agency, US Department of Homeland Security: Washington, DC, USA, 2013. Available online:
https://www.fema.gov/sites/default/files/2020-08/fema_dam-safety_inundation-mapping-flood-risks.pdf (accessed on 25
May 2023).
8. CDSO. Guidelines for Mapping Flood Risk Associated with Dams; CDSO_GUD_DS_05_v1.0; Central Dam Safety Organization, Central
Water Commission, Dam Safety Rehabilitation Directorate: New Delhi, India, 2018. Available online: https://damsafety.cwc.gov.
in/ecm-includes/PDFs/Guidelines_for_Mapping_Flood_Risks_Associated_with_Dams.pdf (accessed on 25 May 2023).
9. Morris, M.W. CADAM Concerted Action on Dam Break Modeling; Final Report SR 571; HR Wallingford: Oxford, UK, 2000; Available
online: https://eprints.hrwallingford.com/447/1/CADAM.pdf (accessed on 25 May 2023).
10. Rizzo, C.; Maranzoni, A.; D’Oria, M. Probabilistic Mapping and Sensitivity Assessment of Dam-Break Flood Hazard. Hydrol. Sci.
J. 2023, 68, 700–718. [CrossRef]
11. Maranzoni, A.; D’Oria, M.; Rizzo, C. Quantitative Flood Hazard Assessment Methods: A Review. J. Flood Risk Manag. 2023, 16,
e12855. [CrossRef]
12. Ferrari, A.; Vacondio, R.; Mignosa, P. High-Resolution 2D Shallow Water Modelling of Dam Failure Floods for Emergency Action
Plans. J. Hydrol. 2023, 618, 129192. [CrossRef]
13. Aureli, F.; Maranzoni, A.; Petaccia, G.; Soares-Frazão, S. Review of Experimental Investigations of Dam-Break Flows over Fixed
Bottom. Water 2023, 15, 1229. [CrossRef]
14. Aureli, F.; Maranzoni, A.; Petaccia, G. Review of Historical Dam-Break Events and Laboratory Tests on Real Topography for the
Validation of Numerical Models. Water 2021, 13, 1968. [CrossRef]
15. Teng, J.; Jakeman, A.J.; Vaze, J.; Croke, B.F.W.; Dutta, D.; Kim, S. Flood Inundation Modelling: A Review of Methods, Recent
Advances and Uncertainty Analysis. Environ. Model. Softw. 2017, 90, 201–216. [CrossRef]
16. De Marchi, G. Sull’Onda di Piena che Seguirebbe al Crollo della Diga di Cancano [On the Dam-Break Wave Resulting from the
Collapse of the Cancano Dam]. L’Energia Elettr. 1945, 22, 157–169. (In Italian)
17. Antunes do Carmo, J.S. Physical Modelling vs. Numerical Modelling: Complementarity and Learning. Preprints 2020, 2020070753.
Available online: https://www.preprints.org/manuscript/202007.0753/v2/download (accessed on 21 July 2023).
18. Mudashiru, R.B.; Sabtu, N.; Abustan, I.; Balogun, W. Flood Hazard Mapping Methods: A Review. J. Hydrol. 2021, 603, 126846.
[CrossRef]
19. Bates, P.D. Flood Inundation Prediction. Annu. Rev. Fluid Mech. 2022, 54, 287–315. [CrossRef]
20. Toro, E.F.; Garcia-Navarro, P. Godunov-Type Methods for Free-Surface Shallow Flows: A Review. J. Hydraul. Res. 2007, 45,
736–751. [CrossRef]
21. Castro-Orgaz, O.; Hager, W.H. Shallow Water Hydraulics; Springer: Cham, Switzerland, 2019.
22. Toro, E.F. Shock-Capturing Methods for Free-Surface Shallow Flows; John Wiley & Sons: Chichester, UK, 2001.
23. Castro-Orgaz, O.; Hager, W.H. Non-Hydrostatic Free Surface Flows; Springer: Cham, Switzerland, 2017.
24. Maranzoni, A.; Tomirotti, M. New Formulation of the Two-Dimensional Steep-Slope Shallow Water Equations. Part I: Theory and
Analysis. Adv. Water Resour. 2022, 166, 104255. [CrossRef]
25. Maranzoni, A.; Tomirotti, M. New Formulation of the Two-Dimensional Steep-Slope Shallow Water Equations. Part II: Numerical
Modeling, Validation, and Application. Adv. Water Resour. 2023, 177, 104403. [CrossRef]
26. Stansby, P.K.; Chegini, A.; Barnes, T.C.D. The Initial Stages of Dam-Break Flow. J. Fluid Mech. 1998, 374, 407–424. [CrossRef]
27. Ozmen-Cagatay, H.; Kocaman, S. Dam-Break Flows During Initial Stage Using SWE and RANS Approaches. J. Hydraul. Res. 2010,
48, 603–611. [CrossRef]
28. Soares-Frazão, S.; Zech, Y. Dam Break in Channels with 90◦ Bend. J. Hydraul. Eng. 2002, 128, 956–968. [CrossRef]
29. Kocaman, S.; Ozmen-Cagatay, H. The Effect of Lateral Channel Contraction on Dam Break Flows: Laboratory Experiment. J.
Hydrol. 2012, 432–433, 145–153. [CrossRef]
30. Khoshkonesh, A.; Nsom, B.; Bahmanpouri, F.; Dehrashid, F.A.; Adeli, A. Numerical Study of the Dynamics and Structure of a
Partial Dam-Break Flow Using the VOF Method. Water Resour. Manage. 2021, 35, 1513–1528. [CrossRef]
Water 2023, 15, 3130 23 of 26

31. Soares-Frazão, S. Experiments of Dam-Break Wave Over a Triangular Bottom Sill. J. Hydraul. Res. 2007, 45 (Suppl. S1), 19–26.
[CrossRef]
32. Ozmen-Cagatay, H.; Kocaman, S.; Guzel, H. Investigation of Dam-Break Flood Waves in a Dry Channel with a Hump. J.
Hydro-Environ. Res. 2014, 8, 304–315. [CrossRef]
33. Soares-Frazão, S.; Zech, Y. Experimental Study of Dam-Break Flow against an Isolated Obstacle. J. Hydraul. Res. 2007, 45
(Suppl. S1), 27–36. [CrossRef]
34. Aureli, F.; Dazzi, S.; Maranzoni, A.; Mignosa, P.; Vacondio, R. Experimental and Numerical Evaluation of the Force Due to the
Impact of a Dam-Break Wave on a Structure. Adv. Water Resour. 2015, 76, 29–42. [CrossRef]
35. Khoshkonesh, A.; Daliri, M.; Riaz, K.; Dehrashid, F.A.; Bahmanpouri, F.; Di Francesco, S. Dam-Break Flow Dynamics over a
Stepped Channel with Vegetation. J. Hydrol. 2022, 613, 128395. [CrossRef]
36. Zhang, T.; Feng, P.; Maksimović, Č.; Bates, P.D. Application of a Three-Dimensional Unstructured-Mesh Finite-Element Flooding
Model and Comparison with Two-Dimensional Approaches. Water Resour. Manage. 2016, 30, 823–841. [CrossRef]
37. Munoz, D.H.; Constantinescu, G. 3-D Dam Break Flow Simulations in Simplified and Complex Domains. Adv. Water Resour. 2020,
137, 103510. [CrossRef]
38. Fernandez-Feria, R. Dam-Break Flow for Arbitrary Slopes of the Bottom. J. Eng. Math. 2006, 54, 319–331. [CrossRef]
39. Castro-Orgaz, O.; Cantero-Chinchilla, F.N. Non-Linear Shallow Water Flow Modelling over Topography with Depth-Averaged
Potential Equations. Environ. Fluid Mech. 2020, 20, 261–291. [CrossRef]
40. Lu, X.; Dong, B.; Zhang, X. A Two-Dimensional Depth-Integrated Non-Hydrostatic Numerical Model for Nearshore Wave
Propagation. Ocean Model. 2015, 96, 187–202. [CrossRef]
41. Cantero-Chinchilla, F.N.; Bergillos, R.J.; Gamero, P.; Castro-Orgaz, O.; Cea, L.; Hager, W.H. Vertically Averaged and Moment
Equations for Dam-Break Wave Modeling: Shallow Water Hypotheses. Water 2020, 12, 3232. [CrossRef]
42. Denlinger, R.P.; Iverson, R.M. Granular Avalanches Across Irregular Three-Dimensional Terrain: 1. Theory and Computation. J.
Geophys. Res. Earth Surf. 2004, 109, F01014. [CrossRef]
43. Denlinger, R.P.; O’Connell, D.R. Computing Nonhydrostatic Shallow-Water Flow over Steep Terrain. J. Hydraul. Eng. 2008, 134,
1590–1602. [CrossRef]
44. Biscarini, C.; Di Francesco, S.; Manciola, P. CFD Modelling Approach for Dam Break Flow Studies. Hydrol. Earth Syst. Sci. 2010,
14, 705–718. [CrossRef]
45. Lane, S.N.; Bradbrook, K.F.; Richards, K.S.; Biron, P.A.; Roy, A.G. The Application of Computational Fluid Dynamics to Natural
River Channels: Three-Dimensional Versus Two-Dimensional Approaches. Geomorphology 1999, 29, 1–20. [CrossRef]
46. Versteeg, H.K.; Malalasekera, W. An Introduction to Computational Fluid Dynamics. The Finite Volume Method; Longman Scientific &
Technical: Harlow, UK, 1995.
47. Hirt, C.W.; Nichols, B.D. Volume of Fluid (VOF) Methods for the Dynamics of Free Boundaries. J. Comput. Phys. 1981, 39, 201–225.
[CrossRef]
48. Osher, S.; Fedkiw, R.P. Level Set Methods: An Overview and Some Recent Results. J. Comput. Phys. 2001, 169, 463–502. [CrossRef]
49. Sussman, M.; Puckett, E.G. A Coupled Level Set and Volume-of-Fluid Method for Computing 3D and Axisymmetric Incompress-
ible Two-Phase Flows. J. Comput. Phys. 2000, 162, 301–337. [CrossRef]
50. Marsooli, R.; Wu, W. 3-D Finite-Volume Model of Dam-Break Flow over Uneven Beds Based on VOF Method. Adv. Water. Resour.
2014, 70, 104–117. [CrossRef]
51. Monaghan, J.J. Simulating Free Surface Flows with SPH. J. Comput. Phys. 1994, 110, 399–406. [CrossRef]
52. Xu, X. An Improved SPH Approach for Simulating 3D Dam-Break Flows with Breaking Waves. Comput. Methods Appl. Mech. Eng.
2016, 311, 723–742. [CrossRef]
53. Xu, X.; Jiang, Y.-L.; Yu, P. SPH Simulations of 3D Dam-Break Flow Against Various Forms of the Obstacle: Toward an Optimal
Design. Ocean Eng. 2021, 229, 108978. [CrossRef]
54. Cleary, P.W.; Prakash, M. Discrete-Element Modelling: Methods and Applications in the Environmental Sciences. Phil. Trans. R.
Soc. Lond. A 2004, 362, 2003–2030. [CrossRef] [PubMed]
55. Wu, J.; Bao, K.; Zhang, H. Research Progress on Dam-Break Floods. In Proceedings of the 2nd IEEE International Conference on
Emergency Management and Management Sciences, Beijing, China, 8–10 August 2011; pp. 334–338. [CrossRef]
56. Ferrari, A.; Fraccarollo, L.; Dumbser, M.; Toro, E.F.; Armanini, A. Three-Dimensional Flow Evolution after a Dam Break. J. Fluid
Mech. 2010, 663, 456–477. [CrossRef]
57. Xie, Z.; Stoesser, T.; Xia, J. Simulation of Three-Dimensional Free-Surface Dam-Break Flows over a Cuboid, Cylinder, and Sphere.
J. Hydraul. Eng. 2021, 147, 6021009. [CrossRef]
58. Issakhov, A.; Imanberdiyeva, M. Numerical Simulation of the Movement of Water Surface of Dam Break Flow by VOF Methods
for Various Obstacles. Int. J. Heat Mass Transf. 2019, 136, 1030–1051. [CrossRef]
59. Issakhov, A.; Borsikbayeva, A. The Impact of a Multilevel Protection Column on the Propagation of a Water Wave and Pressure
Distribution During a Dam Break: Numerical Simulation. J. Hydrol. 2021, 598, 126212. [CrossRef]
60. Luo, P.; Luo, M.; Li, F.; Qi, X.; Huo, A.; Wang, Z.; He, B.; Takara, K.; Nover, D.; Wang, Y. Urban Flood Numerical Simulation:
Research, Methods and Future Perspectives. Environ. Model. Softw. 2022, 156, 105478. [CrossRef]
61. Mignot, E.; Dewals, B. Hydraulic Modelling of Inland Urban Flooding: Recent Advances. J. Hydrol. 2022, 609, 127763. [CrossRef]
Water 2023, 15, 3130 24 of 26

62. Kumar, V.; Sharma, K.V.; Caloiero, T.; Mehta, D.J.; Singh, K. Comprehensive Overview of Flood Modeling Approaches: A Review
of Recent Advances. Hydrology 2023, 10, 141. [CrossRef]
63. Avila-Aceves, E.; Plata-Rocha, W.; Monjandin-Armenta, S.A.; Rangel-Peraza, J.G. Geospatial Modelling of Floods: A Literature
Review. Stoch. Environ. Res. Risk Assess. 2023. [CrossRef]
64. TELEMAC-3D—3D Hydrodynamics. Available online: http://www.opentelemac.org/index.php/presentation?id=18 (accessed
on 27 March 2023).
65. FLOW-3D. Available online: https://www.flow3d.com/products/flow-3d/ (accessed on 27 March 2023).
66. About OpenFOAM. Available online: https://www.openfoam.com/ (accessed on 27 March 2023).
67. DualSPHysics: From Fluid Dynamics to Multiphysics Problems. Available online: https://dual.sphysics.org/ (accessed on 5
April 2023).
68. Amicarelli, A.; Manenti, S.; Albano, R.; Agate, G.; Paggi, M.; Longoni, L.; Mirauda, D.; Ziane, L.; Viccione, G.; Todeschini, S.; et al.
SPHERA v. 9.0.0: A Computational Fluid Dynamics Research Code, Based on the Smoothed Particle Hydrodynamics Mesh-Less
Method. Comput. Phys. Commun. 2020, 250, 107157. [CrossRef]
69. Roubtsova, V.; Kahawita, R. The SPH Technique Applied to Free Surface Flows. Comput. Fluids 2006, 35, 1359–1371. [CrossRef]
70. Cleary, P.W.; Prakash, M.; Rothauge, K. Combining Digital Terrain and Surface Textures with Large-Scale Particle-Based
Computational Models to Predict Dam Collapse and Landslide Events. Int. J. Image Data Fusion 2010, 1, 337–357. [CrossRef]
71. Prakash, M.; Rothauge, K.; Cleary, P.W. Modelling the Impact of Dam Failure Scenarios on Flood Inundation Using SPH. Appl.
Math. Model. 2014, 38, 5515–5534. [CrossRef]
72. Ye, F.; Wang, H.; Ouyang, S.; Tang, X.; Li, Z.; Prakash, M. Spatio-Temporal Analysis and Visualization Using SPH for Dam-Break
and Flood Disasters in a GIS Environment. In Proceedings of the 2012 International Symposium on Geomatics for Integrated
Water Resource Management, Lanzhou, China, 19–21 October 2012. [CrossRef]
73. Cleary, P.W.; Prakash, M.; Mead, S.; Tang, X.; Wang, H.; Ouyang, S. Dynamic Simulation of Dam-Break Scenarios for Risk Analysis
and Disaster Management. Int. J. Image Data Fusion 2012, 3, 333–363. [CrossRef]
74. Lee, E.-S.; Violeau, D.; Issa, R.; Ploix, S. Application of Weakly Compressible and Truly Incompressible SPH to 3-D Water Collapse
in Waterworks. J. Hydraul. Res. 2010, 48 (Suppl. S1), 50–60. [CrossRef]
75. Caboussat, A.; Boyaval, S.; Masserey, A. On the Modeling and Simulation of Non-Hydrostatic Dam Break Flows. Comput. Visual.
Sci. 2011, 14, 401–417. [CrossRef]
76. Vassilevski, Y.V.; Nikitin, K.D.; Olshanskii, M.A.; Terekhov, K.M. CFD Technology for 3D Simulation of Large-Scale Hydrodynamic
Events and Disasters. Russ. J. Numer. Anal. Math. Model. 2012, 27, 399–412. [CrossRef]
77. Vacondio, R.; Mignosa, P.; Pagani, S. 3D SPH Numerical Simulation of the Wave Generated by the Vajont Rockslide. Adv. Water
Res. 2013, 59, 146–156. [CrossRef]
78. Zhainakov, A.Z.; Kurbanaliev, A.Y. Verification of the Open Package OpenFOAM on Dam Break Problems. Thermophys. Aeromech.
2013, 20, 451–461. [CrossRef]
79. Jainakov, A.; Kurbanaliev, A.; Oskonbaev, M. Large-Scale Modeling of Dam Break Induced Flows. In Dam Engineering; Tosun, H.,
Ed.; IntechOpen: London, UK, 2019; pp. 59–72. [CrossRef]
80. Džebo, E.; Žagar, D.; Krzyk, M.; Četina, M.; Petkovšek, G. Different Ways of Defining Wall Shear in Smoothed Particle
Hydrodynamics Simulations of a Dam-Break Wave. J. Hydraul. Res. 2014, 52, 453–464. [CrossRef]
81. Zhou, Z.; Wang, X.; Sun, R.; Ao, X.; Sun, X.; Song, M. Study of the Comprehensive Risk Analysis of Dam-Break Flooding Based on
the Numerical Simulation of Flood Routing. Part II: Model Application and Results. Nat. Hazards 2014, 72, 675–700. [CrossRef]
82. Biscarini, C.; Di Francesco, S.; Ridolfi, E.; Manciola, P. On the Simulation of Floods in a Narrow Bending Valley: The Malpasset
Dam Break Case Study. Water 2016, 8, 545. [CrossRef]
83. TELEMAC Modelling System. 3D Hydrodynamics, TELEMAC-3D Software; Version 7.0, Validation Document; EDF R&D: Paris,
France, 2016; Available online: http://www.opentelemac.org/index.php/component/jdownloads/summary/44-v7p0/1302-
telemac3d-validation-v7p0?Itemid=54 (accessed on 30 March 2023).
84. Amicarelli, A.; Kocak, B.; Sibilla, S.; Grabe, J. A 3D Smoothed Particle Hydrodynamics Model for Erosional Dam-Break Floods.
Int. J. Comput. Fluid Dyn. 2017, 31, 413–434. [CrossRef]
85. Wang, X.; Chen, W.; Zhou, Z.; Zhu, Y.; Wang, C.; Liu, Z. Three-Dimensional Flood Routing of a Dam Break Based on a
High-Precision Digital Model of a Dense Urban Area. Nat. Hazards 2017, 86, 1147–1174. [CrossRef]
86. Wang, K.; Yang, P.; Hudson-Edwards, K.A.; Lyu, W.; Yang, C.; Jing, X. Integration of DSM and SPH to Model Tailings Dam Failure
Run-Out Slurry Routing Across 3D Real Terrain. Water 2018, 10, 1087. [CrossRef]
87. Zhang, T.; Peng, L.; Feng, P. Evaluation of a 3D Unstructured-Mesh Finite Element Model for Dam-Break Floods. Comput. Fluids
2018, 160, 64–77. [CrossRef]
88. Chen, H.-X.; Li, J.; Feng, S.-J.; Gao, H.-Y.; Zhang, D.-M. Simulation of Interactions Between Debris Flow and Check Dams on
Three-Dimensional Terrain. Eng. Geol. 2019, 251, 48–62. [CrossRef]
89. Kurbanaliev, A.I.; Maksutov, A.R.; Obodoeva, G.S.; Oichueva, B.R. Using OpenFOAM Multiphase Solver InterFoam for Large
Scale Modeling. In Proceedings of the 27th World Congress on Engineering and Computer Science, San Francisco, CA, USA, 22–24
October 2019; International Association of Engineers: Hong Kong, China; pp. 366–370. Available online: https://www.iaeng.org/
publication/WCECS2019/WCECS2019_pp366-370.pdf (accessed on 25 August 2023).
Water 2023, 15, 3130 25 of 26

90. Issakhov, A.; Zhandaulet, Y. Numerical Study of Dam Break Waves on Movable Beds for Complex Terrain by Volume of Fluid
Method. Water Resour. Manage. 2020, 34, 463–480. [CrossRef]
91. Wang, K.; Yang, P.; Yu, G.; Yang, C.; Zhu, L. 3D Numerical Modelling of Tailings Dam Breach Run Out Flow over Complex
Terrain: A Multidisciplinary Procedure. Water 2020, 12, 2538. [CrossRef]
92. Yu, D.; Tang, L.; Chen, C. Three-Dimensional Numerical Simulation of Mud Flow from a Tailing Dam Failure Across Complex
Terrain. Nat. Hazards Earth Syst. Sci. 2020, 20, 727–741. [CrossRef]
93. Zhuang, Y.; Yin, Y.; Xing, A.; Jin, K. Combined Numerical Investigation of the Yigong Rock Slide-Debris Avalanche and Subsequent
Dam-Break Flood Propagation in Tibet, China. Landslides 2020, 17, 2217–2229. [CrossRef]
94. Amicarelli, A.; Manenti, S.; Paggi, M. SPH Modelling of Dam-break Floods, with Damage Assessment to Electrical Substations.
Int. J. Comput. Fluid Dyn. 2021, 35, 3–21. [CrossRef]
95. Karam, W.; Khan, F.A.; Alam, M.; Ali, S. Simulation of Dam-Break Flood Wave and Inundation Mapping: A Case Study of
Attabad Lake. Int. J. 2021, 9, 703–714. [CrossRef]
96. Miliani, S.; Montessori, A.; La Rocca, M.; Prestininzi, P. Dam-Break Modeling: LBM as the Way Towards Fully 3D, Large-Scale
Applications. J. Hydraul. Eng. 2021, 147, 4021017. [CrossRef]
97. Ai, C.; Ma, Y.; Ding, W.; Xie, Z.; Dong, G. Three-Dimensional Non-Hydrostatic Model for Dam-Break Flows. Phys. Fluids 2022,
34, 22105. [CrossRef]
98. Issakhov, A.; Borsikbayeva, A.; Abylkassymova, A.; Issakhov, A.; Khikmetov, A. Numerical Modeling of the Dam-Break Flood
over Natural Rivers on Movable Beds. Int. J. Nonlinear Sci. Numer. Simul. 2022. [CrossRef]
99. Yang, Y.; Zhou, X.; Chen, X.; Xie, C. Numerical Simulation of Tailings Flow from Dam Failure over Complex Terrain. Materials
2022, 15, 2288. [CrossRef] [PubMed]
100. Zhuang, Y.; Jin, K.; Cheng, Q.; Xing, A.; Luo, H. Experimental and Numerical Investigations of a Catastrophic Tailings Dam Break
in Daye, Hubei, China. Bull. Eng. Geol. Environ. 2022, 81, 9. [CrossRef]
101. Jiang, H.; Zhao, B.; Dapeng, Z.; Zhu, K. Numerical Simulation of Two-Dimensional Dam Failure and Free-Side Deformation Flow
Studies. Water 2023, 15, 1515. [CrossRef]
102. Oertel, M.; Bung, D.B. Comparison of 2D Dam-Break Waves with VOF and SPH Method. In Proceedings of the 35th IAHR
World Congress, Chengdu, China, 8–13 September 2013; Tsinghua University Press: Beijing, China. Available online: https:
//www.iahr.org/library/infor?pid=14676 (accessed on 24 August 2023).
103. Purbasari, R.J.; Suryanto, A.; Anam, S. Numerical Simulations of Dam-Break Flows by Lattice Boltzmann Method. AIP Conf. Proc.
2021, 2021, 60027. [CrossRef]
104. Maquignon, N.; Smaoui, H.; Sergent, P.; Bader, B. A Simplified and Stable Lattice Boltzmann Shallow Water Model. J. Phys. Conf.
Ser. 2022, 2202, 12055. [CrossRef]
105. LeVeque, R.J. Finite Volume Methods for Hyperbolic Problems; Cambridge University Press: Cambridge, UK, 2002.
106. Dick, E. Introduction to Finite Element Methods in Computational Fluid Dynamics. In Computational Fluid Dynamics, 3rd ed.;
Wendt, J.F., Ed.; Springer: Berlin, Germany, 2009; pp. 235–274.
107. Hervouet, J.-M. Hydrodynamics of Free Surface Flows: Modelling with the Finite Element Method; Wiley: Chichester, UK, 2007.
108. Munoz, D.H.; Constantinescu, G. A Fully 3-D Numerical Model to Predict Flood Wave Propagation and Assess Efficiency of
Flood Protection Measures. Adv. Water Resour. 2018, 122, 148–165. [CrossRef]
109. Rong, Y.; Zhang, T.; Zheng, Y.; Hu, C.; Peng, L.; Feng, P. Three-Dimensional Urban Flood Inundation Simulation Based on Digital
Aerial Photogrammetry. J. Hydrol. 2020, 584, 124308. [CrossRef]
110. Peng, L.; Zhang, T.; Li, J.; Feng, P. Three-Dimensional Numerical Study of Dam-Break Flood Impacting Problem with VOF
Method and Different Turbulence Closures. Water Resour. Manage. 2023, 37, 3875–3895. [CrossRef]
111. Pu, J.H.; Shao, S.; Huang, Y.; Hussain, K. Evaluations of SWEs and SPH Numerical Modelling Techniques for Dam Break Flows.
Eng. Appl. Comput. Fluid Mech. 2013, 7, 544–563. [CrossRef]
112. Issakhov, A.; Zhandaulet, Y.; Nogaeva, A. Numerical Simulation of Dam Break Flow for Various Forms of the Obstacle by VOF
Method. Int. J. Multiph. Flow 2018, 109, 191–206. [CrossRef]
113. Park, I.-R.; Kim, K.-S.; Kim, J.; Van, S.-H. Numerical Investigation of the Effects of Turbulence Intensity on Dam-Break Flows.
Ocean Eng. 2012, 42, 176–187. [CrossRef]
114. Larocque, L.A.; Imran, J.; Chaudhry, M.H. 3D Numerical Simulation of Partial Breach Dam-Break Flow Using the LES and k–
Turbulence Models. J. Hydraul. Res. 2013, 51, 145–157. [CrossRef]
115. Yang, S.; Yang, W.; Qin, S.; Li, Q. Comparative Study on Calculation Methods of Dam-Break Wave. J. Hydraul. Res. 2019, 57,
702–714. [CrossRef]
116. Simsek, O.; Islek, H. 2D and 3D Numerical Simulations of Dam-Break Flow Problem with RANS, DES, and LES. Ocean Eng. 2023,
276, 114298. [CrossRef]
117. Neal, J.; Villanueva, I.; Wright, N.; Willis, T.; Fewtrell, T.; Bates, P. How Much Physical Complexity is Needed to Model Flood
Inundation? Hydrol. Process. 2012, 26, 2264–2282. [CrossRef]
118. Goutal, N. The Malpasset Dam Failure. An Overview and Test Case Definition. In Proceedings of the 4th CADAM Meeting,
Zaragoza, Spain, 18–19 November 1999.
119. Hervouet, J.-M.; Petitjean, A. Malpasset Dam-Break Revisited with Two-Dimensional Computations. J. Hydraul. Res. 1999, 37,
777–788. [CrossRef]
Water 2023, 15, 3130 26 of 26

120. Alcrudo, F.; Mulet, J. Description of the Tous Dam Break Case Study (Spain). J. Hydraul. Res. 2007, 45 (Suppl. S1), 45–57.
[CrossRef]
121. Pilotti, M.; Maranzoni, A.; Tomirotti, M.; Valerio, G. 1923 Gleno Dam Break: Case Study and Numerical Modeling. J. Hydraul.
Eng. 2011, 137, 480–492. [CrossRef]
122. Testa, G.; Zuccalà, D.; Alcrudo, F.; Mulet, J.; Soares-Frazão, S. Flash Flood Flow Experiment in a Simplified Urban District. J.
Hydraul. Res. 2007, 45 (Suppl. S1), 37–44. [CrossRef]
123. Pilotti, M.; Milanesi, L.; Bacchi, V.; Tomirotti, M.; Maranzoni, A. Dam-Break Wave Propagation in Alpine Valley with HEC-RAS
2D: Experimental Cancano Test Case. J. Hydraul. Eng. 2020, 146, 5020003. [CrossRef]
124. Güney, M.S.; Tayfur, G.; Bombar, G.; Elci, S. Distorted Physical Model to Study Sudden Partial Dam Break Flows in an Urban
Area. J. Hydraul. Eng. 2014, 140, 5014006. [CrossRef]
125. Norton, J. An Introduction to Sensitivity Assessment of Simulation Models. Environ. Model. Softw. 2015, 69, 166–174. [CrossRef]
126. Pianosi, F.; Beven, K.; Freer, J.; Hall, J.W.; Rougier, J.; Stephenson, D.B.; Wagener, T. Sensitivity Analysis of Environmental Models:
A Systematic Review with Practical Workflow. Environ. Model. Softw. 2016, 79, 214–232. [CrossRef]
127. Castro-Orgaz, O.; Hager, W.H.; Katopodes, N.D. Variational Models for Nonhydrostatic Free-Surface Flow: A Unified Outlook to
Maritime and Open-Channel Hydraulics Developments. J. Hydraul. Eng. 2023, 149, 4023014. [CrossRef]
128. Yu, D.; Tang, L.; Ye, F.; Chen, C. A Virtual Geographic Environment for Dynamic Simulation and Analysis of Tailings Dam Failure.
Int. J. Digit. Earth 2021, 14, 1194–1212. [CrossRef]
129. Macchione, F.; Costabile, P.; Costanzo, C.; De Santis, R. Moving to 3-D Flood Hazard Maps for Enhancing Risk Communication.
Environ. Model. Softw. 2019, 111, 510–522. [CrossRef]
130. Spero, H.R.; Vazquez-Lopez, I.; Miller, K.; Joshaghani, R.; Cutchin, S.; Enterkine, J. Drones, Virtual Reality, and Modeling:
Communicating Catastrophic Dam Failure. Int. J. Digit. Earth 2022, 15, 585–605. [CrossRef]
131. FLOW-3D Modeling Capabilities. Hybrid Shallow Water/3D Flow. Available online: https://www.flow3d.com/modeling-
capabilities/hybrid-shallow-water-3d-flow/ (accessed on 27 April 2023).

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like