You are on page 1of 17

Designing Experiments to Study Dam Breach

Hydraulic Phenomena
Sílvia Amaral 1; Laura Caldeira 2; Teresa Viseu 3; and Rui M. L. Ferreira 4

Abstract: The success of experimental studies to investigate hydraulics of dam breaching requires an adequate specification of geotechnical
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

parameters and control over its realization in field or in laboratory conditions. Although this is known, there are few studies specifically ad-
dressed to formulate and discuss the requirements for comparable laboratory hydraulic tests to study embankment failure by overtopping. This
study addresses this research gap specifically for homogeneous earthfill dams. It is aimed at proposing and discussing a set of core procedures
and monitoring techniques to be observed in studies of dam breaching by overtopping. A first procedural step involves the theoretical discussion
of similarities between model and prototype erodibility. Further steps involve observations of breached model dams under controlled hydraulic
and geotechnical conditions. Requirements for comparability of hydraulic laboratory practice are described and discussed, and it is argued that
the key geotechnical parameters of control are the soil composition (more specifically, the fraction of fine material) and the compaction con-
ditions (relative compaction and water content). To attain similarity conditions in what concerns model dam erodibility, whose failure modes
should include headcuts and undercutting, a range of reference values for the geotechnical control parameters is suggested and construction/
compaction methods are discussed. DOI: 10.1061/(ASCE)HY.1943-7900.0001678. © 2020 American Society of Civil Engineers.
Author keywords: Hydraulics of dam breaching; Breach discharge; Geotechnical characterization; Dam monitoring.

Introduction take off until the mid-1980s. Early studies explicitly acknowledged
that laboratory research should be considered qualitative (Powledge
Dam failures pose significant flood risks. Such failures can incur et al. 1989a). Difficulties in scaling the model results to the proto-
severe economic and environmental consequences and cause the type size, even in geotechnical centrifuge modeling studies, were
loss of human lives (Wahl 1998; Yochum et al. 2008). Failure stud- promptly identified (Ko et al. 1985; Gilbert and Miller 1991). Never-
ies on embankment dams are of particular interest because they theless, advances were attained in the description of the evolution of
account for three-fourths of the 33,105 large dams worldwide the breach (Simmler and Samet 1982; Pugh 1985; Fujita and Tamura
[earthdams ¼ 88%; rockfill dams ¼ 12% (ICOLD 2003)] and 1987; Powledge et al. 1989b; Fletcher and Gilbert 1992; AlQaser
are, by far, those failing more frequently. A thorough investigation and Ruff 1993) and identification of effective protection measures
of International Commission on Large Dams (ICOLD), United [reviewed by Powledge et al. (1989a) in the United Kingdom and
States Cold Storage (USCOLD), Australian National Commitee United States).
on Large Dams Incorporated (ANCOLD), and other databases Since the late 1990s, several national and international joint
showed that overtopping is responsible for about 48.4% of dam research actions have attested to a renewed interest on embankment
failures up to 1986 (Foster et al. 2000). breach studies, for instance concerted action on dambreak modeling
The study of the hydraulics of dam breaching relied mostly on (CADAM) (Morris 2000), investigation of extreme flood processes
field surveys of breached dams and paleo data (reviewed by and uncertainty (IMPACT) (Morris et al. 2009, 2010), US Depart-
Costa 1985; ASCE 2011). Although some laboratory studies were ment of Agriculture—Agricultural Research Service (USDA-ARS)
performed in the 1960 and 1970s, laboratory investigations did not (e.g., Hanson et al. 2005b; Zhang et al. 2009), US Bureau of Rec-
lamation (USBR) (Wahl and Lentz 2011), and National Research
1
Postdoctoral Researcher, Dept. of Hydraulics and Environment, Institute of Science and Technology for Environment and Agricul-
National Laboratory for Civil Engineering, Avenida do Brasil, 101, Lisbon ture (IRSTEA) (Bonelli et al. 2018). Several other research efforts
1700-066, Portugal (corresponding author). ORCID: https://orcid.org/0000
contributed significantly to the body of empirical data describ-
-0002-3520-3444. Email: samaral@lnec.pt
2
Senior Researcher with Habilitation and Head of LNEC ’s Geothecnics
ing embankment failure by overtopping produced in laboratory con-
Dept., National Laboratory for Civil Engineering, Avenida do Brasil, 101, ditions, such as those of Visser (1998), Coleman et al. (2002),
Lisbon 1700-066, Portugal. Email: laurac@lnec.pt Franca and Almeida (2002), Chinnarasri et al. (2004), Schmocker
3
Assistant Researcher and Head of the Water Resources and Hydraulic and Hager (2010), Orendorff et al. (2013), Zhao et al. (2013a), and
Structures Division, Dept. of Hydraulics and Environment, National Feliciano Cestero et al. (2015), and more recently, Mohamed and
Laboratory for Civil Engineering, Avenida do Brasil, 101, Lisbon 1700-066, El-Ghorab (2019), Ettema and Ng (2016), Rifai et al. (2017), Bento
Portugal. Email: tviseu@lnec.pt et al. (2017), Saghaee et al. (2017), Amaral (2017), and among
4
Associate Professor, Civil Engineering Research and Innovation others.
for Sustainability, Instituto Superior Técnico, Universidade de Lisboa, In these studies, the experimental study of dam breach hydraulics
Avenida Rovisco Pais 1, Lisbon 1049-001, Portugal. Email: ruimferreira@
frequently involved determining the time scales of dam failure,
tecnico.ulisboa.pt
Note. This manuscript was submitted on August 6, 2018; approved quantification of breach flow kinematics, characterization of the in-
on June 25, 2019; published online on January 28, 2020. Discussion terplay between dam erosion mechanisms and flow kinematics, and
period open until June 28, 2020; separate discussions must be submitted quantification of the evolution of dam morphology. Instrumentation
for individual papers. This paper is part of the Journal of Hydraulic to acquire velocity and morphology data varies with the specific
Engineering, © ASCE, ISSN 0733-9429. objectives of the work and with conditions for optical access and

© ASCE 04020014-1 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


intrusiveness of measuring probes, among others. Employing non- embankments have been discussed by Jack (1996), Powledge et al.
intrusive optical methods and image processing methods to deter- (1989b), Zhu et al. (2011), Orendorff (2009), and Schmocker and
mine near-breach flow velocities (Orendorff et al. 2011; Bento et al. Hager (2012). For clay soils (cohesive material), the main erosion
2017; Amaral et al. 2018), breach evolution (Van Emelen et al. mechanisms of the breach is associated with the headcuts that deepen
2015; Bento et al. 2017), or dam morphology (Amaral 2017; and enlarge over time, migrating upstream through the embankment
Ferreira et al. 2018) is a clearly a current trend. This might (Hanson et al. 2005a; AlQaser and Ruff 1993; Zhang et al. 2009).
require laser-based illumination solutions (Van Emelen et al. 2015; For granular soils (noncohesive material), the failure process tends to
Bento et al. 2017) or digital high-speed imaging (Feliciano Cestero develop through a hydraulic progressive/regressive surface erosion
et al. 2015; Zhang et al. 2009) along with more traditional monitor- (Visser 1998; Coleman et al. 2002).
ing solutions for water-table levels (acoustic or resistive probes) and In the presence of impervious elements, such as clay cores or
output and inlet discharges (Zhang et al. 2009). Kinect sensors have upstream blankets (membranes), the erosion starts at the dam
also been applied (Amaral et al. 2018) to monitor breach and dam toe, propagating rapidly up the slope until it reaches the crest, low-
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

morphologic evolution. Other techniques involve the use of tracers ering it and allowing for greater flow rates, which accelerates the
embedded in the dam body (Feliciano Cestero et al. 2015) or (highly erosion process. After the impervious element is exposed, it canti-
intrusive) contact probing (Mohamed and El-Ghorab 2019). levers and sequentially fails in tension (Simmler and Samet 1982).
The range of scales investigated varies from the almost proto- However, there are still open research questions that require more
type scale as those performed under the scope of IMPACT Project complete instrumentation, mostly related to the feedback relations
(Vaskinn et al. 2004; Hanson et al. 2005b) open-air lab, or between the morphological evolution and flow hydraulics through
(Mohamed and El-Ghorab 2019) large-scale tests to small-scale the body of the embankment. These are particularly relevant in stud-
embankments (Schmocker and Hager 2010; Feliciano Cestero et al. ies aimed at clarifying processes occurring in embankments built
2015). Both dams and dykes have been investigated, the latter nor- with real soils or realistic sediment mixtures, which necessarily in-
mally in conditions of frontal overtopping (Fletcher and Gilbert volve dealing with cohesive sediment (Amaral et al. 2018). A fun-
1992; Visser 1998; Pickert et al. 2011; Feliciano Cestero et al. damental premise of this study is that further advances in the
2015) but also in conditions of lateral flow (Wahl and Lentz 2011; knowledge on dam and dyke breach hydraulics require comparabil-
Rifai et al. 2017). Most studies involved homogeneous embank- ity among studies with standardization of procedures and an ad-
ments, but zoned (Fletcher and Gilbert 1992; Morris et al. 2009, equate specification of geotechnical parameters and control over
2010) and embankments with impervious membranes (Simmler its realization in field or laboratory conditions. It is noticed that some
and Samet 1982; Franca and Almeida 2002) were also studied. previous studies have had evident care in the characterization of the
In the case of homogeneous embankments, different types of geotechnical properties of the embankment (Fletcher and Gilbert
material have been investigated, including (1) uniform sand, for 1992; Zhao et al. 2013a; Feliciano Cestero et al. 2015; Amaral
which pore pressure may introduce apparent cohesion (Simmler 2017). Some others mention relevant quantities but not a full geo-
and Samet 1982; Visser 1998; Coleman et al. 2002; Pickert technical characterization (Morris et al. 2009, 2010). In fact, the
et al. 2004; Schmocker and Hager 2010; Orendorff et al. 2013); classification of cohesive embankment means different things in dif-
(2) rip-rap (Franca and Almeida 2002); (3) poorly sorted noncohe- ferent studies (Hanson et al. 2005a; Feliciano Cestero et al. 2015;
sive material (Chinnarasri et al. 2004); (4) poorly sorted material Zhao et al. 2013b), which is surely due to the lack of precise def-
including clay sizes conferring the mixture cohesive properties (Ko initions in hydraulic studies dealing with geotechnical processes.
et al. 1985; Hanson et al. 2005a; Morris et al. 2009, 2010; Zhao There are other aspects about model design and construction
et al. 2013b; Feliciano Cestero et al. 2015); and (5) actual soil with techniques that require attention, including the implementation
cohesive properties (Hanson et al. 2005a; Morris et al. 2009, 2010; of bottom drainage for seepage with toe drains (Orendorff
Bento et al. 2017; Amaral 2017). In the case of cohesive material, 2009; Walder et al. 2015; Al-Riffai 2014), compaction protocol
studies have addressed the influence of the grain-size distribution (Feliciano Cestero et al. 2015; Zhang et al. 2009), or monitoring
(Feliciano Cestero et al. 2015) as well as water content and com- strategy. No studies have specifically sought to present and dis-
paction energy (Morris et al. 2009, 2010; Zhang et al. 2009). cuss the core procedures and monitoring practices for a geotechni-
The type of overtopping has largely differed, with some studies cally well-posed experimental hydraulic study of embankment
focusing on full-crest overtopping (Simmler and Samet 1982; Ko failure by overtopping.
et al. 1985; Powledge et al. 1989a; Schmocker and Hager 2010; This paper addresses this research gap specifically for homo-
Chinnarasri et al. 2004), with others aimed at studying the mech- geneous earthfill dams built with realistic soils. It is aimed at for-
anisms of breach development employing initial notches carved in mulating and discussing a set of core procedures and monitoring
the embankment crest (Visser 1998; Coleman et al. 2002; Morris techniques to be observed in hydraulic dam breach studies to ensure
et al. 2009, 2010; Orendorff et al. 2009, 2013; Zhao et al. 2015; comparability among experimental tests of similar size. To meet
Feliciano Cestero et al. 2015; Mohamed and El-Ghorab 2019; this objective, (1) theoretical arguments about similarity of dam
Bento et al. 2017; Amaral 2017). Other studies looked at (1) the erodibility are formulated, (2) the procedure to select a range of
effect of soil strength on scour depth at abutments formed of erod- soils and compaction techniques is described, and (3) the results
ible, compacted earthen embankments built with both noncohesive of the laboratory tests featuring embankments built with different
and cohesive soils (Ettema and Ng 2016); and (2) the quantification soils and compacted under different conditions are analyzed.
of the erosion resistance by establishing a novel setup device that
allowed the overflowing rate to increase (Bonelli et al. 2018). The
location of this notch has also been investigated (Morris et al. 2009, Scaling Dam Erodibility: Problem Statement
2010; Zhao et al. 2013a).
An essentially consistent description of embankment breaching The failure by overtopping of homogeneous earthfill dams involves
by overtopping has emerged from these studies. There is a consensus hydraulic and geotechnical phenomena. To obtain similarity in the
that different failure mechanisms may be observed for embankments mechanical behavior of the breached embankment, its characteris-
composed of different soils (Singh 1996), and differences between tics cannot be defined by hydraulic similarity considerations only.
overtopping failure in clayed (cohesive) and granular (noncohesive) The forces dominating the hydraulic processes are gravity and

© ASCE 04020014-2 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


inertial forces, which call for Froude similarity. However, the erod- The sediment transport rate in the model (assuming a constant
ibility of reduced earthfill dam models depends on hydraulic forces α, the same fluid in the model and prototype, and the same gravity
but also on soil resistance properties, including apparent or actual field, i.e., tests not carried out in a centrifuge) is
mineralogical cohesion, which do not obey Froude similarity.  m
pffiffiffiffiffiffiffiffiffi τM τ yM
Furthermore, the hydraulic forces involved in erosive processes qsM ¼ αðdM gdM Þ − ð4Þ
may require other types of similarity, namely Reynolds similarity ðρsM − ρÞgdM ðρsM − ρÞgdM
(e.g., Gilbert and Miller 1991), which is impossible to attain using
the same fluid in the model and the prototype and under the same From Froude similarity, one expects
gravity field.
It is well-known from observations in large-scale models and qsM ¼ λ3=2
L qsP ð5Þ
prototypes (e.g., Mohamed et al. 2002; Hanson et al. 2005b;
Morris et al. 2010) that earthfill dams subjected to overtopping As argued previously, the threshold stress is a property of the
soil—cohesion and microfrictional contacts—but the hydraulic ap-
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

feature hydraulic erosion and mass detachment episodes. If these


modes of failure are to be observed in the scaled models, extra plied stress is a property of fluid flow. If the model is built with the
similarity considerations and a distinct geometrical scaling for soil same soil (or finer) than the prototype and the bulk density is the
particles must be considered. same, it would be difficult to achieve positive values of τ − τ y ow-
The conditions to approach similarity of erodibility are dis- ing to the smaller size of the model. In other words, it would be
cussed next. The term bulk erodibility shall refer to all processes impossible to guarantee similar erodibility.
that result in loss of dam sediment mass. Hydraulic erodibility To understand how much the soil resistance would have to be
refers to processes associated to (continuous) hydraulic erosion. lowered in the model, one should first define the applied hydraulic
Erodibility of an embankment is not an intrinsic property of its soil, stress
such as other parameters used in geotechnical design such as τ ∝ fρU 2 ð6Þ
strength, permeability, and porosity, among others.
Hydraulic erodibility is defined for a particular load condition, where U = characteristic flow velocity; and f = Darcy-Weisbach
and it is expressed by two concepts: soil erosion resistance (or easi- friction factor. The latter is a function of the Reynolds number
ness of erosion) and the rate of soil erosion. The soil erosion re-
sistance has been quantified by several researchers (Arulanandan Uδ
Re ¼ ð7Þ
and Perry 1983; Arulanandan et al. 1975; Wan and Fell 2004) ν
by a critical or threshold shear stress, τ y (ML−1 T−2 ). This is de-
where δ = thickness of the boundary layer of the flow over the dam
fined as the value of the hydraulic shear stress required to initiate
body; and ν = fluid’s kinematic viscosity. Assuming that the fluid is
erosion during a hydraulic flow. Erosion of the soil will occur if the
the same in the model and in the prototype, and given that
shear stress applied by flowing water, τ (ML−1 T−2 ), exceeds the
λU ¼ λ1=2
L , then
critical (or threshold) shear stress of the soil, τ y .
The soil erosion rate can be expressed is the rate of mass re- ReM ¼ λ3=2
L ReP ð8Þ
moval per unit area. Using Exner’s equation (and assuming a
one-dimensional flow), it can be expressed (Exner 1925) i.e., the Reynolds number in the model is much smaller than the
Reynolds number in the prototype. Therefore, the friction factor
∂Zb ∂q is larger in the model. Anomalous Froude scaling is thus obtained
ð1 − pÞ ¼− s ð1Þ
∂t ∂x for the hydraulic stresses
where p = soil porosity; zb = elevation of the dam body subjected to fM
hydraulic erosion (L); and qs = sediment transport rate per unit τ M ¼ λL τ P ð9Þ
fP
width (L2 T−1 ).
Major variables that influence qs include (1) soil and sediment From Colebrook-White’s formula for open-channel flows (Graf
properties, including bulk sediment density [ρs (ML−3 )], cohesion 1971; Bousmar and Zech 1999) and considering that the boundary
[σ0 (ML−1 T−2 )], apparent mineralogical origin, and grain-size dis- layer over the breach crest is poorly developed (Bento et al. 2017),
tribution d (L); (2) fluid properties [density, ρ (ML−3 ) and viscosity, one has that fM =f P is less than 10 and most likely less than 5.
μ (ML−1 T−1 )]; and (3) applied hydraulic forces per unit area (τ ) Given that the geometric scale based on the height of the prototype
and gravity, g (MLT−2 ). In general and model dams may be much less than 1=10, one concludes that
τ M is indeed smaller than τ P , even if not as small as if strict Froude
qs ¼ qs ðρs ; d; σ0 ; ρ; μ; τ ; gÞ ð2Þ
similarity would hold.
Buckingham’s Pi theorem (Brand, 1957) and phenomenological As an example, assume that a 0.5-m-tall model is simulating
considerations lead to the following formula: the processes seen in a 10-m-tall prototype. Assume that ReP ¼
 m 2 × 105 . The geometric scale would be λL ¼ 1=20 and ReM ¼
qs τ τy 2,236. Then f M =fP ¼ 3.3 and τ M ¼ 0.167τ P . In order to maintain
pffiffiffiffiffi ¼ α − ð3Þ the imbalance τ M − τ yM equal to that of the prototype, the soil’s
d gd ðρs − ρÞgd ðρs − ρÞgd
resistance property τ yM should be reduced to 16.7% of its value
where α = parameter that may be a weak function of the grain-size in the prototype. Evidently, this can only be achieved by reducing
distribution and bulk sediment density, where in particular, α would soil bulk density or decreasing the percentage of fines to reduce
increase with decreasing grain size and decreasing bulk soil den- cohesion. More importantly, there is no objective or correct value
sity; and τ y = threshold stress, below which there is no soil erosion. of bulk density or correct grain-size distribution if the aim of
Eq. (3) may not be general, but it expresses the most relevant the study is not modeling a specific (prototype) dam. If the aim
traits of hydraulic erosion: it only exists if τ > τ y , it increases non- is characterizing and quantifying processes, the geometric scale
linearly with the excess of the hydraulic action, and it increases is arbitrary, and the choice of grain-size distribution and bulk den-
with the reduction of soil bulk density, the same imbalance τ − τ y . sity may be governed by other considerations, namely availability

© ASCE 04020014-3 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


of material and reproduction of large-scale features of observed grain-size distribution to ensure proportional erodibility or reduce
geotechnical behavior. bulk density by applying a lower compaction energy.
Once it is guaranteed that both hydraulic stresses and soil resis- From the preceding discussion, it is clear that the model’s grain-
tance obey the same scaling, one may cast Eq. (4) as follows: size distribution cannot be scaled in accordance to the ratio of dam
pffiffiffiffiffiffiffiffiffi heights. The resulting dam model would have mechanical proper-
ðd gdM Þ ties very different from the prototype, ultimately leading to a bulk
qsM ¼ α M  ðΔτ M Þm ð10Þ
ðρsM gdM Þm erodibility much lower than that required to exhibit the failure
mechanisms of prototypes.
where Δτ M ¼ τ M − τ yM ; and ρsM ¼ ρsM − ρM . Assuming that the Two remarks should be made regarding the foregoing theoreti-
geometric scale of soil particle diameter may be different from the cal analysis:
geometric scale and that there is also a degree of freedom between • The soil particle scaling in Eq. (13) is not a general proposal; it
ρsM and ρsP , Eq. (10) becomes is an example of application of the similarity arguments devel-
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

  oped for hydraulic erodibility. That this example leads coarser


qSP ðλd Þ3=2 fP m
qsM ¼ ðλL Þm ð11Þ sediment in the model is incidental, but it is not absurd. Eqs. (12)
ðλρs Þm ðλd Þm fM and (13) are an attempt to introduce rationality in the process of
choosing the sediment mixture in the model, but the actual scal-
Eq. (11) shows that if the exponent m is 3/2 [the exponent of ing can be more complex if one takes into account all processes
Meyer, Peter, and Müller formula (Meyer-Peter and Müller 1948), involved in bulk erodibility.
for instance], there is no possibility to achieve similarity in erodibility • If the ultimate goal is to study hydraulics of dam failure by over-
given that the Reynolds numbers would be always different. In that topping, the soil should be chosen so that model bulk erodibility
case, changing the gravity field would be necessary. If m is not 3/2, allows for the development of the particular erosion modes of
from Eqs. (5) and (11), one may obtain the geometrical scaling of interest, including mass detachment episodes. Possibly, the most
soil material that would ensure similarity of erodibility as follows: efficient way to achieve this is to first test a model built with soil
  m scaled in accordance with Eq. (12), then perform preliminary
f m−3=2
λd ¼ λL M ðλρs Þm ð12Þ tests with varying soil composition and degree of compaction
fP to make sure that the model does exhibit all modes of erosion
expected in prototypes. It should thus be clear that the preceding
Further pursuing the last example, if the soil in the model is analysis does not open the possibility that model results are ex-
compacted to 80% of the bulk density of the prototype, and if actly scaled to prototype sizes. Rather, the analysis opens the
m ¼ 2, then possibility of hydraulic testing in the model under some degree
of similarity in what concerns bulk erodibility.
dM ¼ 3.8dP ð13Þ
Once a range of soil and compaction energy are established, it is
i.e., a coarser grain-size distribution would be needed in the model. necessary to (1) determine the sufficient set of geotechnical design
Further research is needed to specify parameter m. parameters and its values to ensure that the embankment has the
Again, without an imposed value of λL (to model a specific required erodibility conditions; (2) set the building protocol for
dam), the grain-size distribution should be chosen to guarantee the model embankment, e.g., number and thickness of soil layers,
the macroscopic geotechnical behavior, that is bulk erodibility. as well as which compaction method, plus weight and plate size of
A trial-and-error procedure, such as that described in the next the compactor; and (3) adopt a monitoring procedure to ensure that
section, may be appropriate. the design parameters are attained. The following sections address
In addition to hydraulic erodibility, several processes involving these issues.
discontinuous mass failure may be present and determine dam bulk
erodibility. For instance, underscouring creates conditions for mass
detachments episodes laterally (Morris et al. 2010; Amaral et al. Preliminary Identification of Types of Soil
2018) or by creating conditions for toppling (Fei et al. 2014; and Definition of Compacting Techniques
Ashourian et al. 2018). Other discrete failure mechanisms of geo-
technical origin may be possible, but the focus here is on processes Preliminary Selection of Types of Soil
that depend on underscouring of dams as the headcut reaches a for Hydraulic Testing
fixed bottom. Its evolution rate depends on the mean velocity
and turbulence of the flow generated by the spread of the jetlike A set of six homogeneous earthfill dams, about 0.45 m high and
flow released from the breach as it hits the fixed bottom 1.5 m wide, were conducted to failure by overtopping under the
(Feliciano Cestero et al. 2015). The mean velocity should obey scope of this study (Fig. 1). For this purpose, the soil of several
Froude similarity because it depends essentially on gravity and homogeneous earthfill dams was first analyzed to define a range
height of dam. Turbulence intensities should be higher in the proto- of realistic prototype grain-size distribution curves [gray zone in
type, but the proper scaling should be based on the Reynolds num- Fig. 2(a)]. The upper limit of this range is a silty sand (SM) or
ber with the integral length scale as a geometric variable. Because a clayed sand (SC) [Fig. 2(a)] and the lower limit is a lean clay
the latter should obey geometric similarity based on the dam height (CL) [Fig. 2(a)], according to the unified soil classification system
(Chachereau and Chanson 2011), Eq. (8) should be valid for this (USCS) [ASTM D2487 (2011)]. The lower limit has necessarily to
type of erosion. Turbulence intensities should be a monotonous in- be lean clay—low plasticity—because in earthfill dams, fat clays—
creasing function of the Reynolds number, although this is not highly plastic—must be excluded due to their highly expansive and
firmly established in this case; the closest applicable research con- deformable behavior.
cerns turbulence in hydraulic jumps (e.g., Mignot and Cienfuegos Several samples of sand and silty and clayed soils were mixed
2010). It is thus legitimate to expect that underscouring erosion and tested along with samples of natural soils, whose grain-size
in the model would be smaller than that occurring under strict distribution curves were determined by sieving and sedimentation
Froude similarity. Again, it would necessary to employ a coarser analysis [LNEC E 196 (LNEC 1966)].

© ASCE 04020014-4 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Homogeneous embankment dam model: (a) downstream slope; (b) top view with the crest width; and (c) upstream slope.

Fig. 2. (a) Range of grain-size distribution curves that fall under the type of those concerning Portuguese homogeneous earthfill dams; and (b) grain-
size distribution curves of clay-sand mixture [clayed sand (SC)] (dashed line), natural soil [silty sand (SM)] (solid line), and target range of grading
curves (gray area).

Table 1. Tested soils; grading properties, Atterberg limits, density, and permeability coefficient
Atterberg limits (%)
Sample wP wL PI G K (m=s)
Clayed sand (SC) D50 ¼ 0.23 mm; 37% fines; 61% sand; 2% gravel 16.5 34.8 18.2 2.67 6 × 10−10
Silty sand (SM) D50 ¼ 0.31 mm; 27% fines; 66% sand; 7% gravel — 17.5 Nonplastic 2.64 1 × 10−7
Note: wP = plastic limit; wL = liquid limit; PI = plasticity index; G = particle density; and K = permeability coefficient.

Two different soils, a clay-sand mixture (1:2 clay to sand) and a point of compaction for the natural soil are, respectively, 4.0 and
natural soil (from a quarry extraction deposit) were selected for 36.7 N=m2 , which qualitatively correspond to a soil with a mod-
hydraulic testing. The clay-sand mixture was artificially processed erately rapid rate of progression of internal erosion. These values
with a cement mixer. Their grain-size distributions can be seen in were obtained from the hole erosion test (HET) as I HET and τ HET
Fig. 2(b). Their grading properties and the corresponding Atterberg (Wan and Fell 2004; Wahl and Erdogan 2008). Although the
limits [wP , wL , and plasticity index (PI)], soil particle density (G), goal of HET is to evaluate the internal erodibility of the soil, it
and the permeability coefficient (k) can be found in Table 1. These is relatable with the jet erosion test, which evaluates the superficial
parameters were obtained according to NP-143 (LNEC 1969), erodibility (Regazzoni et al. 2008). Thus, under appropriate com-
NP-83 (LNEC 1965), and ASTM D7664 (ASTM 2018). paction conditions, it was envisaged that this natural soil would
The clay-sand mixture and the natural soil are a clayed exibit similarity in terms of concerns erodibility conditions.
sand (SC) and a silty sand (SM) (ASTM D2487) with 37% and
27% of fines content (Table 1), respectively. Both soils have a
Compaction Methods
percentage of fines similar to those usually adopted in actual
dams (12%–50%). Both have a sufficiently low permeability for Three different compaction methods were tested (Fig. 3): (1) using
the purpose of impounding water (ksilty sand ðSMÞ ¼ 10−7 m=s and a lawn roller filled with water without vibration; (2) by percussion,
kclayed sand ðSCÞ ¼ 10−10 m=s, as indicated in Table 1). using a metallic hand tamper; and (3) by vibration, using a low-
The natural soil is slightly coarser than the upper limit of the weight vibratory plate [PC1010 Euroshatal (Gersthofen, Germany)
grain-size distributions of Fig. 2(a). Although this is not necessary, low-weight compactor model (50 kg) with plate size ¼ 30 × 43 cm
the size of finer grain sizes of the natural soil approximately obey and vibration frequency ¼ 100 Hz].
Eq. (13) if one considers an average grain-size distribution curve The first method (compaction using a lawn roller filled with
located between the upper and lower limits present in Fig. 2(a) water) revealed itself to be inappropriate because the lack of vibra-
The index of erosion rate and the critical shear stress at the optimum tion combined with the low weight of the roller resulted in lifts

© ASCE 04020014-5 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Compaction of the experimental embankment dams tested using (a) lawn roller; (b) metallic hand tamper; and (c) vibratory plate. (Images by
authors.)

compacted only close to the surface (first 2–3 cm). The collected Fig. 5 shows that relative compactions within range (80%–90%)
samples disaggregated after extraction [Figs. 4(a and b)]. can be obtained for the natural soil with both tested methods:
The other two compaction methods were evaluated by asses- (1) percussional for all the tested compaction loads (4, 8, and
sing the physical conditions of the soil after compaction for three 12 blows) at the optimum water content (woptimum ); and (2) vibra-
different compaction loads, materialized in terms of number of tional for one or three passes of the vibratory plate at dry conditions
blows (4, 8, and 12) [Figs. 4(c and d)] or passages (1, 3, and 6) (w < woptimum ). A single pass would not assure a homogenous em-
[Figs. 4(e–g)], respectively, for the compaction by percussion or bankment. Six passes with the vibratory plate resulted in too high
vibration. The standard Proctor compaction curve (natural soil), relative compactions (96%). The compaction by percussion also
together with the dry unit weights (γ d ) and water content (w) cor- shown to be inefficient in terms of ensuring homogeneity when
responding to the evaluated compaction loads for each compaction compared with the vibratory compaction.
method, percussion and vibratory, are presented in Figs. 5(a and b), The clayed sand soil was compacted by percussion using a
respectively. metallic hand tamper by applying a number of blows compatible

Fig. 4. Trial lifts for testing the compaction methods: (a) sample collection; (b) samples showing disaggregation; (c) bands with different compaction
loads; (d) sample locations; (e) trial lift; and (f and g) sand replacement method to determine the unit weight of the compacted soil. (Images by
authors.)

© ASCE 04020014-6 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


Fig. 5. Physical conditions of the soil after compaction: (a) by percussion (4, 8, and 12 blows); and (b) by vibration (1, 3, and 6 passes).
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

with the reference compaction conditions from the standard The characterization of the tests is given in Table 2, including type
Proctor test. of soil, water content scenario, and the corresponding properties.
The model earthfill dams were approximately 0.45–0.48 m tall
and 1.50 m wide and were built in a 31.50-m-long and 6.60-m-wide
Hydraulic Testing to Define Suitable Dam Material, hydraulic channel (Fig. 6) at the Portuguese National Laboratory
Geotechnical Parameters, and Monitoring for Civil Engineering (LNEC). The model dam’s crest width
Procedures was 0.17 m, and the face slopes were 1V∶2H (upstream) and
1V∶2.5H (downstream). All model dams where equipped with a
downstream toe drain for seepage control (Fig. 1). The total volume
Layout of Laboratory Tests
of the reservoir was approximately 60 m3 to ensure negligible
In field tests, such as those performed under the scope of CADAM, kinetic head upstream the dam site.
IMPACT, and FLOODsite projects (Mohamed et al. 2002; Morris Each test involved the following steps:
et al. 2009, 2010), it was observed that the failure process of real 1. Preparatory work:
embankment dams (prototypes) subjected to overtopping involves a. prepare/mix soils according to goal proportions of each type
undercutting, underscouring leading to the appearance of lateral of material,
cavities, and consequent detachment of masses of soil from the b. perform the standard Proctor test (determine both optimum
dam body. Before studying hydraulic phenomena associated with soil water content wopt and corresponding maximum dry unit
these erosion modes, it was necessary to test the types of soil pre- weight γ max
d ),
viously selected as well as the building techniques to achieve model c. construct the earthfill dam by adopting a multiple lift system
dams that possess an erodibility similar to that envisaged in and implementing a compaction effort correspondent to pre-
prototypes. designated water content and dry unit weight, and
Six different tests were performed combining the two types of 2. Dam failure test:
soil described in the “Preliminary Selection of Types of Soil for a. partial fill of the reservoir to establish steady-state seepage
Hydraulic Testing” section, namely clayed sand (SC) and natural conditions,
soil (silty sand), and several values of relative compaction and b. determination of a common zero for all instrumentation, for
water content in relation to the results of the standard Proctor test. instance water-level probes and video monitoring equipment,

Table 2. Main characteristics of the experimental tests: main outcomes


Soil characteristics Hydraulic control
Tested water
Test Soil content scenario γ d (kN=m3 ) w (%) RC (%) Δw (%) tf Qmax (L=s) Observed erosion patterns
T0 Clayed — 18.8 12.0 a
— — 5 h 30 min 2.3 Hydraulic erosion at an ≈
sand constant rate and undercutting
(SC) formation
Compaction by percussion with a metallic hand tamper
T1 Silty wtested ∈ woptimum  2% 16.00 10.5 81 1.2 1 min 12 s 340 Hydraulic erosion; headcut formation,
T2 sand 17.30 10.6 87 1.2 17 min 40 s 253 and sudden sidewall collapses;
(SM) underscouring especially in T2
Compaction using a vibratory plate compactor of low weight (50 kg)
Tc 1 Silty wtested < woptimum − 2% 17.46 6.2 88 −3.3 3 min 14 s 331 Hydraulic erosion; headcut
Tc 2 sand 17.53 6.6 87 −2.9 4 min 48 s 370 formation and sudden sidewall
Tc 3 (SM) 16.83 7.4 85 −2.0 —a —a collapses induced by underscouring
in test Tc3 the regressive erosion
was very marked
Note: γ d = dry unit weight; w = water content; RC = relative compaction referred to the standard Proctor test; Δw = deviation from woptimum ; tf = failure time;
and Qmax = maximum effluent flow.
a
Tc 3 was stopped during the breach development phase, and the failure time and maximum effluent flow were not characterized.

© ASCE 04020014-7 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Plan view of the laboratory channel (dimensions in meters).

c. acquisition of calibration images for digital imaging acquisi- increase and when it reaches the peak value, maximum effluent
tion methods, flow, Qmax , observed erosion modes, and dam morphology.
d. carving a pilot channel, if the location of the breach is to be The evolution of the erosion process was recorded with high
controlled, speed and regular [25 frames per second (fps)] digital video from
e. fill the remaining reservoir up to the crest level and activate a overhead and frontal cameras, thus allowing for the identification
trigger to start all acquisition devices, of erosion modes. A motion sensor, Kinect (Microsoft, Redmond,
f. enforce the prespecified plan for the evolution of the water Washington), was used for a nonintrusive assessment of breach mor-
level in the reservoir (for instance a constant water level at phology and for three-dimensional (3D) embankment reconstruction.
a prespecified elevation), The latter was done only for the final dam geometry. Either the breach
g. consider the test finished when relevant phenomena has been discharge was monitored in all tests by directly measuring the breach
observed, and area and flow velocity or by computing the mass balance in the res-
h. record all data acquired with digital systems and convert to ervoir and the outflow through a triangular weir installed downstream
digital and store any analogue data [examples of operation the model dam (details have been given by Bento et al. 2017).
have been given by Orendorff et al. (2013), Zhang et al.
(2009), and Bento et al. (2017)].
Relevant variables to characterize the hydraulic behavior of the Description of the Clay Sand Test T0
model included the failure time, tf , which corresponds to the time Test T0 consisted of testing a model dam built with the clayed sand
interval between the instants when the breach effluent flow starts to (SC) soil (Fig. 7). This soil is gap-graded and thus inadequate for

Fig. 7. Preliminary test: temporal evolution of the embankment dam failure (breaching failure). (Images by authors.)

© ASCE 04020014-8 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


Fig. 8. Test 1 (T1) erosion processes: (a) headcut formation (with steps and pools; and (b) sidewall collapse in the left side of the breach. (Images by
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

authors.)

actual prototype dams. The negative aspects of this soil, such as (2) to avoid the mixing procedure and instead select a natural soil
segregation during application (stratification) and suffusion ten- fitting the goal range; (3) to decrease the embankment’s relative
dency during the reservoir filling (or the experimental test itself), compaction (to increase the embankment erodibility); and (4) to
were mitigated by the quality of the mixer procedure (with a cement vary the water content among tests.
mixer) and by the reduced dimensions of the model dam, respec-
tively. The mixing procedure was appropriate to ensure homo- Description of Silty Sand (SM) Tests T1, T2, T c 1,
geneity, but is laborious and very time-consuming. T c 2, and T c 3
Compaction was aimed at reproducing the optimal conditions of
the standard Proctor test. It was performed by percussion using a
Compaction Scenarios: Relative Compaction
metallic hand tamper. After compaction (for the optimum water
and Water Content
content), and given that the percentage of material retained by The reference compaction conditions from the standard Proctor
No. 200 sieve was about 63.0%, the soil mixture showed an evident test (LNEC E 197) for this soil were first determined: γ max ¼
d
agglutinate behavior and high critical shear resistance (low erodi- 19.9 kN=m3 and woptimum ¼ 9.4%. Then, several compaction con-
bility). The test was stopped 5.5 h after its initiation. During the ditions (relative compaction and water content) were tested. The
failure process, only hydraulic erosion was observed and at low following protocol was adopted:
rates. There was evident headcut erosion in the downstream slope 1. Adopt relative compactions (RC) varying between 80% and
with minor steps and pools events (this is clear in Fig. 7 at 2:00). 90% of the standard Proctor reference values.
No mass detachment episodes were observed. 2. Adopt two different scenarios of water content deviations from
The results of this test were mainly of qualitative nature and the reference value ðwoptimum Þ:
served mostly to improve the experimental procedure of further • w ∈ woptimum  2% is within the standard practice reference
tests. These are the major practical conclusions: interval for actual dam building; and
• The soil grain-size distribution is not only inappropriate for • w < woptimum − 2% is below the usual lower limit in order to
prototype dams (gap-graded soil), but it is also inadequate reduce shear resistance [Eqs. (5)–(10) are discussed in the
for models given that it leads to a disproportionally large shear “Scaling Dam Erodibility: Problem Statement” section].
resistance.
• Compaction by percussion was revealed to be appropriate Role of Relative Compaction: Comparison between Tests T1
for this type of soil, but the adoption of the reference values from and T2
the standard Proctor test contributed for the excess of embank- Tests T1 and T2, compacted by percussion, evaluated the sensitiv-
ment resistance, hence not meeting the erodibility conditions ity of the breaching process to the relative compaction and hence
required for these tests (discussed in the “Scaling Dam Erodibil- were conducted in equivalent water content conditions.
ity: Problem Statement” section). The compaction loads corresponded to 4 and 12 metallic hand
These conclusions led to the following decisions for the follow- tamper blows, respectively. Bulk densities were within the relative
ing earthfill dam breach tests, performed with the silty-sand (SM) compaction goal (80%–90%), and water contents within the usual
of natural origin: (1) to adopt a less plastic and well-graded soil; reference interval w ∈ woptimum  2% (Table 2).

Fig. 9. Test 2 (T2) erosion processes: (a) erosion cavity created by underscouring; and (b) sidewall collapse caused by the underscouring. (Images by
authors.)

© ASCE 04020014-9 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


The earthfill dam of Test T1 was revealed to be highly erodible, three passes with the vibratory plate. Both tests were conducted
which resulted in an extremely fast erosion of the dam’s material with nearly the same relative compactions of the preliminary Test
(1 min 12 s), attaining a maximum breach effluent flow of 340 L=s T2 but in drier conditions [RC ∈ ð85% − 87%Þ and w < woptimum −
(summarized in Table 2). In addition, headcuts formed almost 2% as summarized in Table 2]. The comparison among these tests
instantaneously, representing a steps and pools pattern along the was aimed at evaluating the sensitivity of the breaching process to
downstream face [Fig. 8(a)]. This observation is in line with the the water content.
stair-step headcuts in the dam downstream face associated with The combination of both low compaction and dry state of the
the first stages of erosion of embankments composed by soils with embankments resulted in highly erodible embankments because
high fines content (with a cohesive mechanical behavior) first both conditions lead to a diminishing of shear resistance. As in-
described by Ralston (1987). Several sidewall collapses were also dicated in Table 2, the failure times corresponding to Tests Tc 1
observed during this fast process. and Tc 2 (3 min 14 s and 4 min 48 s, respectively) are much
In Test T2, the embankment dam exhibited a relative compac- shorter than that of Test T2 (17 min 40 s). The difference of
about 4% in the water contents between the former (Tc 1 and
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

tion higher than that of Test T1, maintaining similar water content
conditions [RC ¼ 87% and w ¼ 1.2% (Table 2)]. As expected, the Tc 2) and the latter (T2) was translated into failure times of about
higher resistance to superficial erosion imposed by the higher rel- one-quarter of that of Test T2. Probably, this difference was fur-
ative compaction resulted in a higher failure time and lower breach ther accentuated by the fact that Tests Tc 1, Tc 2, and T2 had
effluent flow [17 min 40 s and 253 L=s, respectively (Table 2)]. water contents in different sides of the reference curve of the
It was concluded that embankments with similar water contents standard Proctor test (dry and wet sides, respectively). Fig. 10
erode faster and display a higher peak breach discharge for lower shows visual evidence of this increased erodibility of Tests Tc 1
relative compaction conditions. In terms of the erosion patterns, and Tc 2.
hydraulic and headcut erosion were present, as well as sidewall col-
lapses [Fig. 9(b)]. In this case, some of the masses of soil detached Validation (Test Tc 3)
from the dam body occurred by underscouring [Fig. 9(a)]. The embankment of Test Tc 3 was compacted by vibration with
three passes with the vibratory plate. This corresponded to nearly
Role of Water Content and Higher Relative Compaction: the same relative compactions of Test T2 but in drier conditions
Comparison among Tests T2, Tc 1, and Tc 2 [RC ¼ 85% and Δw ¼ −2% (Table 2)]. Relatively to Tests Tc 1
The embankments of Tests Tc 1 and Tc 2 were compacted by and Tc 2, the water content was slightly larger but the compaction
vibration, using as reference a compaction load corresponding to ratio was lower. Analyzing the behavior of the previous tests, it was

Fig. 10. Breach temporal evolution in Tests T2, Tc1, and Tc2, where t1 corresponds to the instant when the breach effluent flow started to increase,
and t4 corresponds to the time instant when the maximum breach flow is attained (peak flow). Failure time ðtfÞ ¼ t4 − t1. (Images by Sílvia Amaral.)

© ASCE 04020014-10 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. Breaching process behavior observed in the homogeneous embankment dam failure test showing stages of the failure during Test Tc3:
(a) images from the frontal HD camera; and (b) images from the top view camera. (Images by authors.)

© ASCE 04020014-11 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


envisaged that such soil and compaction conditions would result in Key Procedures and Monitoring Practices
a failure process with strong hydraulic erosion, visible underscour- for Dam Breaching Tests
ing, and significant mass detachment episodes. In this sense, this
was a validation test. Criteria for Soil Selection and Geotechnical Parameters
Expecting underscouring, a Kinect sensor was deployed to of the Compacted Soil
survey the breach morphology and final breach morphology. For
this purpose, a full 3D scan of the dam body was performed at To characterize the soil of embankments built for dam breaching
an intermediate stage of breach development; the test was finalized tests, the following geotechnical parameters, which characterize
before complete failure. the soil as a geological material and describe it in its compacted
The characterization of the several failure stages observed in form, were considered the indispensable minima:
the earthfill dam of Test Tc 3 is represented in Fig. 11 and described • grading properties (grain-size distribution curve);
next. • compaction curve from the standard Proctor test, particularly
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

Initially, the flow passes through the pilot channel and reaches the optimal values of relative compaction (RC) and water con-
the downstream toe of the dam with a small flow width (Stage 1 in tent (woptimum );
Fig. 11). The erosion process begins when the tractive shear stress • Atterberg limits (wP , wL , and PI);
induced by the overtopping flow exceeds the critical shear stress of • density of soil particles (G);
the embankment material and the downstream slope of the embank- • permeability coefficient (k);
ment starts eroding, mainly by hydraulic erosion. Then, the breach- • index of erosion rate (I) at the optimum point of compac-
ing process initiates with the formation of small steps and pools on tion; and
the downstream slope (a stair step pattern). The amplitude of these • critical shear stress (τ ) at the optimum point of compaction.
increases with time, developing into the formation of one or two The latter two are relevant in the case of clay soils or soils with a
well-defined headcuts (Stage 2 in Fig. 11). As the erosion process cohesive structural behavior. The hole erosion test (Wan and Fell
continues to evolve upstream by regressive erosion (Stages 3 and 4 2004; Wahl and Erdogan 2008) and jet erosion test [Hanson and
in Fig. 11), the headcuts enlarge and occasionally mass detachment Cook 2004; ASTM D5852 (ASTM 1995); Wahl 2016] are two po-
episodes occur. These are most likely by underscouring, i.e., the tential methods for evaluating the progression of erosion in this
erosion cavity created as the jetlike issued by the breach hits type of soils that can be related through empirical expressions.
and spreads on the fixed floor (Fig. 12). As erosion advances, A complete comparison between the two has been given by Regaz-
the cavities enlarge, the turbulence of the effluent flow increases, zoni et al. (2008).
and the mass detachment episodes become more frequent, leading Geotechnical parameters of the soil and compaction conditions
to further enlargement of the erosion cavities. This a nonlinear of the embankment must be unambiguously characterized to ensure
self-feeding process that plays a role in determining the breach comparability of overtopping experiments with homogeneous
failure time. earthfill model dams with similar dimensions. In terms of the se-
When the underscouring cavity starts propagating below and lection of the soil, first, a percentage of fines between 12% and 50%
upstream the dam crest, the breaching process accelerates. The is suggested to ensure that the embankment presents a cohesive
breach enlarges faster and progresses further upstream because structural behavior (silt or clay soils can be chosen). Percentages
mass detachment episodes directly contribute to enlarging the re- above this value will lead to an agglutinate behavior with a dispro-
gion where the subcritical–supercritical flow transition occurs portionally high critical shear resistance [discussed following
(Stages 5–7 in Fig. 11). At this stage, there is an evident increase Eq. (5) in the “Scaling Dam Erodibility: Problem Statement” sec-
in the positive slope of the rising limp of the breach effluent hydro- tion]. Second, well-graded soils are suggested because gap-graded
graph, as well as its corresponding crest, which moves upstream. soils are likely to segregate during application and have a suffusion
The failure mechanisms observed in the experimental tests per- tendency.
formed as part this research, and in particular, in Test Tc 3, are com- To ensure the similarity conditions in terms of hydraulic erod-
patible with the failure behavior of large-scale embankments ibility, the grain-size distribution curve should be scaled in accor-
composed of clayed soils described by Zhang et al. (2009), Hanson dance to Eq. (12) to get an order of magnitude of the model soil. In
et al. (2005b), Hahn et al. (2000), and AlQaser and Ruff (1993). At any case, a reduction in the grain-size distribution curve should
this stage, one might consider that the geotechnical requirements only be implemented if the smaller dimension of the experimental
that guarantee similarity conditions in terms of dam erodibility embankment is smaller than 6–8 times the soil maximum diameter,
are known and unambiguously specified in terms of the traditional D100 [ASTM D2850 (2007)]. In case this, only the gravel fraction
geotechnical parameters featured in Tables 1 and 2. of the soil should be scaled. This can be carried out by adopting the

Fig. 12. Sidewall collapse from the left side of the breach: (a) frontal view of the falling mass; and (b) 3D reconstruction of the morphology of the
failed earth dam after the sidewall collapse (point cloud after postprocessing the Kinect sensor data). (Images by authors.)

© ASCE 04020014-12 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


scalp and replace equation studied by Lin et al. (2001). In most lifts. For the size of the earthfill dams of this study, lifts of
cases, the soil of the prototype can be used in the experimental tests 0.05–0.08 m thickness showed a good compromise between
(Amaral 2017). construction time and adequate compaction (uniform com-
paction throughout the thickness of each lift). The material
Construction of Embankments and Monitoring of each lift must be sprayed with water and compacted to
Procedures ensure uniformity. Compaction by vibration leads to higher
degrees of homogeneity of the earthfill dams than a compac-
To increase the erodibility of the experimental embankments it is tion by percussion and hence should be preferred to the latter
necessary to reduce the relative compaction and to adjust its water to reduce uncertainty. Furthermore, for failure purposes, ex-
content. Embankments will display different physical properties perimental embankment dams with relative compactions
depending on which side of the Proctor reference curve they have varying between 80% and 90% should be adopted to increase
been compacted. A summary of the construction and monitoring the embankment’s bulk erodibility, as a form of compen-
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

procedures emerging as best practice in this work are listed as sating for the reduced value of external actions (velocity
follows: field in the facility) relative to prototype values (“Scaling
1. Use RCs below the standard Proctor reference values. In any Dam Erodibility: Problem Statement” and “Compaction
case, a sensitivity analysis to the RC must be performed pre- Methods” sections). This reference interval for compaction
viously, and the actual RC values must reflect a trade-off is good trade-off between erodibility and realistic geotechni-
between embankment erodibility and realistic geotechnical cal behavior (Fig. 13).
behavior. • Step 2: control the compaction during construction by gath-
2. Create different scenarios of water content deviations from ering samples of compacted soil in each lift and conduct-
the reference value determined from the standard Proctor test ing laboratory analysis involving evaluation of the dry unit
ðwoptimum Þ. Select the actual value based on a sensitivity analysis. weight (γ d ), water content (w), and relative compaction (RC)
3. Evaluate the compaction method, particularly the appropriate- referred to as the standard Proctor test. This can be per-
ness of percussional versus vibrational equipment, bearing in formed by using a sand replacement method, by applying
mind the type of soil and the dimensions of the dam model. a nuclear gauge testing in each lift, or by using other equiva-
4. Develop an embankment construction protocol that covers the lent methods.
following main issues: • Step 3: spray the surface material of the embankment with
• Step 1: adopt a multiple lift or layered system with a thick- water and cover it with a plastic sleeve to prevent variations
ness compatible with a uniform compaction throughout the of humidity while the soil is stored awaiting construction.

Fig. 13. Experimental procedure for the construction of the homogeneous embankment dams. (Images by Sílvia Amaral.)

© ASCE 04020014-13 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


This is required because usually, a few days are necessary to required bulk erodibility conditions; (2) set the building protocol
setup the instrumentation. for the model embankment (in terms of number of soil layers, thick-
• Step 4: perform the final cuts to attain the desired geometry nesses, and compaction method); and (3) decide a monitoring pro-
in the embankments (on the day of the failure test) and im- cedure to ensure that the design parameters are attained. The
plement the toe drain. The most efficient way to construct the particular contributions of this research were based on a set of
embankment is to build it larger than the desired geometry experimental tests on failure by overtopping of homogeneous em-
and then cut it until the desired slopes and crest width are bankment dams focused on sandy soils with high fines content
attained. This procedure will remove the superficial parts [silty sand (SM) and clayey sand (SC) according to the USCS].
of the embankments and those more exposed that have lost These tests encompassed the variation of the type of soil (fines con-
humidity, and thus safeguard the homogeneity of humidity in tent), compaction method (percussion versus vibratory), and soil
the entire dam body. Regarding the toe drain implementa- compaction conditions (relative compaction and water content).
tion, its goal is to assure that internal moisture conditions The influence of classical geotechnical parameters in the breaching
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

are compatible with the functionality of the embankment, process was generically evaluated by monitoring the failure time,
i.e., the internal phreatic line has to be established and con- maximum effluent flow, and erosion patterns.
verge to the dam toe. If the embankment dams have no toe The procedure adopted in the definition of the experimental
drain, the headcut erosion that might appear during the fail- earthfill dams, encompassing the soil selection process and the con-
ure by overtopping is unrealistically large. struction and compaction method, was revealed to be appropriate
• Step 5: carve a pilot channel in the middle of the dam crest to for dam breach experiments because the overtopping breaching
control the area of concentrated flow and the erosion at the processes usually observed in prototypes of homogeneous embank-
beginning of the overtopping process. ment dams were properly reproduced.
The tests carried also contributed to a better understanding of the
influence of the relative compaction and water content of the earth-
Conclusions fill dams in the breaching process. It was observed that earthfill
dams with similar water contents erode faster and present higher
This work addressed the issue of formulating and discussing the peak discharges (breach effluent flow) for lower relative compac-
requirements for comparable laboratory hydraulic tests of embank- tion conditions, i.e., respond in an inversely proportional manner.
ments subjected overtopping, particularly models of homogeneous In addition, low water contents translate into faster erosion,
earthfill dams. Some core procedures and monitoring practices that i.e., drier embankments are more erodible.
can be used to design comparable dam breach tests (hydraulic and
geotechnical) were proposed.
The first challenge concerned the theoretical discussion of Data Availability Statement
the possibility of similarity between model and prototype erodibil-
ity. It was found that the geometrical scaling for the model Some or all data, models, or code generated or used during the
soil should include a correction for simple geometric dam height study are available from the corresponding author by request.
scaling. To attain a model hydraulic erodibility similar to proto-
type erodibility and taking in consideration processes beyond
simple Froude scaling, one arrives at the correction factor Acknowledgments
ðf M =fP Þm=½m−ð3=2Þ ðλρs Þm in Eq. (12). This correction factor im-
The implementation of the methods herein presented was possible
plies that not only must the grain-size distribution must be coarser
due to the support of the Sector of Modelling and Construction,
than that obtained by simple geometric scaling but also the com-
Scientific Instrumentation Centre and Geotechnics Department
paction should be different from that attained in the prototype.
of the Portuguese National Laboratory for Civil Engineering.
Considerations about mass detachment episodes may change the
The work was partially funded under the COMPETE program
actual correction factor but not the key point that model grain-size
(FEDER) and National Funds through FCT Project PTDC/ECI-
distribution cannot be scaled in accordance to the ratio of dam
EGC/31618/2017. The first author is thankful for the financial sup-
heights if the aim is to have the same bulk erodibility in the model
port through the Ph.D. scholarship of FCT, No. SFRH/BD/47694/
and prototype.
2008. All authors are indebted to the three anonymous reviewers
The theoretical arguments developed introduce rationality in the
and the associate editor for the discussions raised in the reviewing
process of choosing the dam model soil and open the possibility of
process that contributed to a stronger manuscript.
hydraulic testing under some degree of similarity in terms of bulk
erodibility. Again, it should be emphasized that this research con-
cerns the conditions of possibility of hydraulic testing; the objective Notation
is to study details of flow hydraulics in model dams that exhibit
failure modes similar to those found in prototypes. In other words, The following symbols are used in this paper:
this analysis does not provide exact scaling laws to generalize for d = grain-size distribution (L);
prototype sizes measurements taken in models, particularly in f = Darcy-Weisbach friction factor;
terms of variables expressing geotechnical behavior. It is thus G = density of soil particles (ML−3 );
implicit that full geotechnical characterization of a breaching
g = gravity (MLT−2 );
dam, including phenomenological advances in failure modes,
can possibly be achieved only at prototype scale. I HET = index of erosion rate (HET);
Once a first criterion for model soil grain size and compaction is K = soil permeability (LT−1 ]);
achieved, further progress involved observations of breached model p = soil porosity;
dams under controlled hydraulic and geotechnical conditions. The Qmax = maximum effluent flow (L3 T);
aims were to (1) determine the sufficient set of geotechnical design qs = sediment transport rate per unit width (L2 T−1 );
parameters and their values to ensure that the embankment has the tf = failure time (T);

© ASCE 04020014-14 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


U = characteristic flow velocity (LT−1 ); Bonelli, S., S. Nicaise, G. Charrier, N. Chaouch, F. Byron, and Y.
w = water content (%); Grémeaux. 2018. “Quantifying the erosion resistance of dikes
with the overflowing simulator.” In Proc., Protections 2018: Int. Conf.
wL = liquid limit (Atterberg limits) (%); on Protection against Overtopping. Wallingford, UK: HR Wallingford.
woptimum = optimum water content (%); Bousmar, D., and Y. Zech. 1999. “Momentum transfer for practical
wP = plastic limit (Atterberg limits) (%); flow computation in compound channels.” J. Hydraul. Eng. 125 (7):
zb = elevation of the dam body subjected to hydraulic 696–706. https://doi.org/10.1061/(ASCE)0733-9429(1999)125:7(696).
erosion (L); Brand, L. 1957. “The Pi theorem of dimensional analysis.” Arch. Ration.
α = weak function of the grain-size distribution and of bulk Mech. Anal. 1 (1): 35–45.
sediment density; Chachereau, Y., and H. Chanson. 2011. “Free-surface fluctuations and tur-
bulence in hydraulic jumps.” Exp. Therm. Fluid Sci. 35 (6): 896–909.
γ d = dry unit weights (ML−2 T−2 );
https://doi.org/10.1016/j.expthermflusci.2011.01.009.
γ max
d = maximum dry unit weight (ML−2 T−2 ); Chinnarasri, C., S. Jirakitlerd, and S. Wongwises. 2004. “Embankment dam
δ = thickness of the boundary layer of the flow over the
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

breach and its outflow characteristics.” Civ. Eng. Environ. Syst. 21 (4):
dam body (L); 247–264. https://doi.org/10.1080/10286600412331328622.
λL = length scale; Coleman, S. E., D. P. Andrews, and M. G. Webby. 2002. “Overtopping
λU = velocity scale; breaching of noncohesive homogeneous embankments.” J. Hydraul.
μ = viscosity (ML−1 T−1 ); Eng. 128 (9): 829–838. https://doi.org/10.1061/(ASCE)0733-9429
(2002)128:9(829).
ν = fluid’s kinematic viscosity (ML−1 T−1 );
Costa, J. E. 1985. Floods from dam failures. Reston, VA: USGS.
ρ = density (ML−3 ); Ettema, R., and K. W. Ng. 2016. “Insights from laboratory experiments
ρs = bulk sediment density (ML−3 ); on the failure of earthen embankments at bridge-waterway abutments.”
σ0 = cohesion (ML−1 T−2 ); In Proc., Protections 2016: 2nd Int. Seminar on Dam Protection
τ = shear stress applied by hydraulic forces (ML−1 T−2 ); Against Overtopping. Fort Collins, CO: Colorado State Univ.
τ HET = critical shear stress (HET) (ML−1 T−2 ); and Exner, F. M. 1925. “Uber die wechselwirkung zwischen wasser und ge-
schiebe in flussen.” Akad. Wiss. Wien Math. Naturwiss. Kl. 134 (2a):
τ y = threshold shear stress of the soil (ML−1 T−2 ).
165–204.
Fei, J., H. Yuan, and Q. Zhang. 2014. “Dam breaching models and soil
mechanics analysis.” In Soil behavior and geomechanics, 699–708.
References Reston, VA: ASCE.
Feliciano Cestero, J. A., J. Imran, and M. H. Chaudhry. 2015. “Experimen-
AlQaser, G., and J. F. Ruff. 1993. “Progressive failure of an overtopped tal investigation of the effects of soil properties on levee breach by over-
embankment.” In Hydraulic engineering, 1957–1962. Reston, VA: topping.” J. Hydraul. Eng. 141 (4): 04014085. https://doi.org/10.1061
ASCE. /(ASCE)HY.1943-7900.0000964.
Al-Riffai, M. 2014. “Experimental study of breach mechanics in Ferreira, R. L., S. Amaral, A. M. Ricardo, and A. M. Bento. 2018. “Assess-
overtopped noncohesive earthen embankments.” Ph.D. thesis, Dept. ing the mass flux from a breached model dam with laser imaging and
of Civil Engineering, Univ. of Ottawa. Large-Scale PIV.” In Proc., 19th Int. Symp. on Applications of Laser
Amaral, S. 2017. “Experimental characterization of the failure by overtop- and Imaging Techniques to Fluid Mechanics. Lisbon, Portugal: Center
ping of embankment dams.” Ph.D. thesis, Instituto Superior Técnico, for Innovation, Technology and Policy Research.
Universidade de Lisboa. Fletcher, B. P., and P. A. Gilbert. 1992. Center Hill Fuseplug Spill-
Amaral, S., T. Alvarez, T. Viseu, and R. M. L. Ferreira. 2018. “Image analy- way Caney Fork River, Tennessee: Hydraulic model investigation.
sis detection applied to dam breach experiments.” In Proc., 5th IAHR Vicksburg, MS: Army Engineer Waterways Experiment Station,
Europe Congress—New Challenges in Hydraulic Research and
Hydraulics Laboratory.
Engineering. Bologna, Italy: Maccaferri.
Foster, M., R. Fell, and M. Spannagle. 2000. “The statistics of embankment
Arulanandan, K. 1975. “Fundamental aspects of erosion of cohesive soils.”
dam failures and accidents.” Can. Geotech. J. 37 (5): 1000–1024.
J. Hydraul. Div. 101 (5): 635–639.
https://doi.org/10.1139/t00-030.
Arulanandan, K., and E. B. Perry. 1983. “Erosion in relation to filter design
Franca, M. J., and A. B. Almeida. 2002. “Experimental tests on rockfill
criteria in earth dams.” J. Geotech. Eng. 109 (5): 682–698. https://doi
dam breaching process.” In Proc., IAHR-Int. Symp. on Hydraulic
.org/10.1061/(ASCE)0733-9410(1983)109:5(682).
and Hydrological Aspects of Reliability and Safety Assessment of
ASCE Task Committee on Dam/Levee Breaching. 2011. “Earthen embank-
Hydraulic Structures, 29–31. St. Petersburg, Russia: Vniig Im. B.ye.
ment breaching.” J. Hydraul. Eng. 137 (12): 1549–1564. https://doi.org
/10.1061/(ASCE)HY.1943-7900.0000498. Vedeneyeva.
Ashourian, M., M. Shafai-Bejestan, and H. Babazadeh. 2018. “Investiga- Fujita, Y., and T. Tamura. 1987. “Enlargement of breaches in flood levees
tion of headcut erosion in cohesive soils.” Water Res. 45 (1): 69–78. on alluvial plains.” Nat. Disaster Sci. 9 (1): 37–60.
ASTM. 1995. Standard test method for erodibility determination of soil in Gilbert, P. A., and S. P. Miller. 1991. A study of embankment performance
the field or in the laboratory by the jet index method. ASTM D5852. during overtopping. Vicksburg, MS: Army Engineer Waterways
West Conshohocken, PA: ASTM. Experiment Station, Geotechnical Laboratory.
ASTM. 2007. Standard test method for unconsolidated-undrained Graf, W. H. 1971. Hydraulics of sediment transport. New York: McGraw-
triaxial compression test on cohesive soils. ASTM D2850. West Hill.
Conshohocken, PA: ASTM. Hahn, W., G. J. Hanson, and K. R. Cook. 2000. “Breach morphology
ASTM. 2011. Standard practice for classification of soils for engineering observations of embankment overtopping tests.” In Proc., Joint Conf.
purposes (unified soil classification system). ASTM D2487. West on Water Resources Engineering and Water Resources Planning and
Conshohocken, PA: ASTM. Management, 1–10. Reston, VA: Environmental and Water Resources
ASTM. 2018. Standard test method for measurement of hydraulic conduc- Institute of ASCE.
tivity of unsaturated soils. ASTM D7664. West Conshohocken, PA: Hanson, G. J., and K. R. Cook. 2004. “Apparatus, test procedures, and ana-
ASTM. lytical methods to measure soil erodibility in situ.” Appl. Eng. Agric.
Bento, A. M., S. Amaral, T. Viseu, R. Cardoso, and R. M. L. Ferreira. 2017. 20 (4): 455. https://doi.org/10.13031/2013.16492.
“Direct estimate of the breach hydrograph of an overtopped earth dam.” Hanson, G. J., K. R. Cook, and S. L. Hunt. 2005a. “Physical modeling of
J. Hydraul. Eng. 143 (6): 06017004. https://doi.org/10.1061/(ASCE) overtopping erosion and breach formation of cohesive embankments.”
HY.1943-7900.0001294. Trans. ASAE 48 (5): 1783–1794. https://doi.org/10.13031/2013.20012.

© ASCE 04020014-15 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


Hanson, G. J., M. Morris, K. Vaskinn, D. Temple, S. Hunt, and M. Hassan. Powledge, G. R., D. C. Ralston, P. Miller, Y. H. Chen, P. E. Clopper, and
2005b. “Research activities on the erosion mechanics of overtopped D. M. Temple. 1989b. “Mechanics of overflow erosion on embank-
embankment dams.” ASDSO J. Dam Saf. 3 (1): 4–16. ments. II: Hydraulic and design considerations.” J. Hydraul. Eng.
ICOLD (International Commission on Large Dams). 2003. Bulletin on risk 115 (8): 1056–1075. https://doi.org/10.1061/(ASCE)0733-9429(1989)
assessment in dam safety management. Paris: ICOLD. 115:8(1056).
Jack, R. 1996. “The mechanics of embankment failure due to overtopping Pugh, C. A. 1985. Hydraulic model studies of fuse plug embankments.
flow.” M.Sc. thesis, Dept. of Civil and Resource Engineering, Univ. of Denver: US Bureau of Reclamation.
Auckland. Ralston, D. C. 1987. “Mechanics of embankment erosion during overflow.”
Ko, H. Y., R. J. Dunn, and T. Hollingsworth. 1985. Study of embankment In Hydraulic engineering, 733–738. Reston, VA: ASCE.
performance during overtopping-prototype modelling and dimensional Regazzoni, P. L., D. Marot, J. R. Courivaud, G. Hanson, and T. Wahl. 2008.
verification. Vicksburg, MS: USACE. “Soil erodibility: A comparison between the jet erosion test and the hole
Lin, D.-F., M. K. Chang, and H. L. Luo. 2001. “Study of scalp-and-replace erosion test.” In Proc., Inaugural Int. Conf. of the Engineering Mechan-
equation in compaction specification for soil with high gravel content.” ics Institute, 1–7. Reston, VA: ASCE.
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

Transp. Res. Rec. 1772 (1): 55–61. https://doi.org/10.3141/1772-07. Rifai, I., S. Erpicum, P. Archambeau, D. Violeau, M. Pirotton, K. El Kadi
LNEC (Laboratório Nacional de Engenharia Civil). 1965. Solos— Abderrezzak, and B. Dewals. 2017. “Overtopping induced failure of
Determinação da densidade das partículas. LNEC NP-83. Lisboa, noncohesive, homogeneous fluvial dikes.” Water Resour. Res.
Portugal: LNEC. 53 (4): 3373–3386. https://doi.org/10.1002/2016WR020053.
LNEC (Laboratório Nacional de Engenharia Civil). 1966. Solos—Análise Saghaee, G., A. A. Mousa, and M. A. Meguid. 2017. “Plausible
granulométrica. LNEC E 196. Lisboa, Portugal: LNEC. failure mechanisms of wildlife-damaged earth levees: Insights from
LNEC (Laboratório Nacional de Engenharia Civil). 1969. Solos— centrifuge modeling and numerical analysis.” Can. Geotech. J. 54 (10):
Determinação de limites de consistência. LNEC NP-143. Lisboa, 1496–1508. https://doi.org/10.1139/cgj-2016-0484.
Portugal: LNEC. Schmocker, L., and W. H. Hager. 2010. “Modelling dike breaching due to
Meyer-Peter, E., and R. Müller. 1948. “Formulas for bed-load transport.” In overtopping.” J. Hydraul. Res. 47 (5): 585–597. https://doi.org/10.3826
Proc., IAHSR 2nd Meeting, Stockholm, Appendix 2. Madrid, Spain: /jhr.2009.3586.
International Association for Hydro-Environment Engineering and Schmocker, L., and W. H. Hager. 2012. “Plane dike-breach due to over-
Research. topping: Effects of sediment, dike height and discharge.” J. Hydraul.
Mignot, E., and R. Cienfuegos. 2010. “Energy dissipation and turbulent Res. 50 (6): 576–586. https://doi.org/10.1080/00221686.2012.713034.
production in weak hydraulic jumps.” J. Hydraul. Eng. 136 (2): Simmler, H., and L. Samet. 1982. “Dam failure from overtopping studied
116–121. https://doi.org/10.1061/(ASCE)HY.1943-7900.0000124. on a hydraulic model.” In Proc., 14th Int. Congress on Large Dams. Rio
Mohamed, M. A. A., P. G. Samuels, M. W. Morris, and G. S. Ghataora. de Janeiro, Brazil: International Commission on Large Dam.
2002. “Improving the accuracy of prediction of breach formation Singh, V. 1996. Vol. 17 of Dam breach modeling technology. New York:
through embankment dams and flood embankments.” In Proc., Int. Springer.
Conf. on Fluvial Hydraulics—River Flow, edited by D. Bousmar Van Emelen, S., Y. Zech, and S. Soares-Frazao. 2015. “Formation de brèches
and Y. Zech. Lisse, Netherlands: Swets and Zeitlinger. dans les digues de sable: modélisation de l'interaction eau-sol.” La Houille
Mohamed, M. M. A., and E. A. S. El-Ghorab. 2019. “Investigating scale Blanche 6 (Dec): 21–28. https://doi.org/10.1051/lhb/20150064.
effects on breach evolution of overtopped sand embankments.” Water Vaskinn, K. A., A. Lovell, K. Hoeg, M. W. Morris, G. J. Hanson, and M. A.
Sci. 30 (2): 84–95. https://doi.org/10.1016/j.wsj.2016.10.003. A. M. Hassan. 2004. “IMPACT: Physical modelling of breach forma-
Morris, M. W. 2000. “CADAM: A European concerted action project on tion: Large scale field tests.” In Proc., Association of State Dam Safety
dam-break modelling.” In Proc., Biennial Conf. of the BDS, 42–53. Officials: Dam Safety Conf. Phoenix: Association of State Dam Safety
London: Thomas Telford. Officials.
Morris, M. W., M. Hassan, and K. A. Vaskinn. 2010. “Breach formation: Visser, P. J. 1998. “Breach erosion in sand-dikes.” In Vol. 3 of Proc.,
Field test and laboratory experiments.” Supplement, J. Hydraul. Res. Int. Conf. on Coastal Engineering, edited by B. L. Edge, 3516–
45 (S1): 9–17. https://doi.org/10.1080/00221686.2007.9521828. 3528. Reston, VA: ASCE.
Morris, M. W., A. Kortenhaus, P. J. Visser, and M. Hassan. 2009. Breach- Wahl, T. L. 1998. Prediction of embankment dam breach parameters—
ing processes: A state of the art review. FLOODsite Rep. T06-06-03. A literature review and needs assessment. Dam Safety Rep No.
Wallingford, UK: FLOODsite Consortium. DSO-98-004. Denver: US Dept. of the Interior, Bureau of Reclamation.
Orendorff, B. 2009. “An experimental study of embankment dam breach- Wahl, T. L. 2016. “The submerged jet erosion test: Past-present-future.”
ing.” Ph.D. thesis, Dept. of Civil Engineering, Univ. of Ottawa. In Proc., USSD Int. Symp. on the Mechanics of Internal Erosion for
Orendorff, B., M. Al-Riffai, I. Nistor, and C. D. Rennie. 2013. “Breach Dams and Levees. Westminster, CO: United States Society on Dams.
outflow characteristics of non-cohesive embankment dams subject to Wahl, T. L., and Z. Erdogan. 2008. Determining erosion indices of cohesive
blast.” Can. J. Civ. Eng. 40 (3): 243–253. https://doi.org/10.1139/cjce soils with the hole erosion test and jet erosion test. Denver: US Dept. of
-2012-0303. Interior, Bureau of Reclamation, Technical Service Center.
Orendorff, B., M. Al-Riffai, C. D. Rennie, I. Nistor, and P. St-Germain. 2009. Wahl, T. L., and D. J. Lentz. 2011. Physical hydraulic modeling of canal
“The effects of initial breach geometry on dam breach morphology.” In breaches. Hydraulic Laboratory Rep. No. HL-2011-09. Denver: US
Proc., 33rd IAHR Congress, 9–14. Madrid, Spain: International Asso- Bureau of Reclamation.
ciation for Hydro-Environment Engineering and Research. Walder, J. S., R. M. Iverson, J. W. Godt, M. Logan, and S. A. Solovitz.
Orendorff, B., C. D. Rennie, and I. Nistor. 2011. “Using PTV through an 2015. “Controls on the breach geometry and flood hydrograph during
embankment breach channel.” J. Hydro-Environ. Res. 5 (4): 277–287. overtopping of noncohesive earthen dams.” Water Resour. Res. 51 (8):
https://doi.org/10.1016/j.jher.2010.12.003. 6701–6724. https://doi.org/10.1002/2014WR016620.
Pickert, G., G. H. Jirka, A. Bieberstein, and J. Brauns. 2004. “Soil/water Wan, C. F., and R. Fell. 2004. “Laboratory tests on the rate of piping
interaction during the breaching process of overtopped embankments.” erosion of soils in embankment dams.” Geotech. Test. J. 27 (3):
In Proc., 2nd Int. Conf. on River Flow, 903–910. Boca Raton, FL: CRC 295–303. https://doi.org/10.1520/GTJ11903.
Press. Yochum, S. E., L. A. Goertz, and P. H. Jones. 2008. “Case study of the big
Pickert, G., V. Weitbrecht, and A. Bieberstein. 2011. “Breaching of over- bay dam failure: Accuracy and comparison of breach predictions.”
topped river embankments controlled by apparent cohesion.” J. Hydraul. J. Hydraul. Eng. 134 (9): 1285–1293. https://doi.org/10.1061/(ASCE)
Res. 49 (2): 143–156. https://doi.org/10.1080/00221686.2011.552468. 0733-9429(2008)134:9(1285).
Powledge, G. R., D. C. Ralston, P. Miller, Y. H. Chen, P. E. Clopper, and Zhang, J., Y. Li, G. Xuan, X. Wang, and J. Li. 2009. “Overtopping breach-
D. M. Temple. 1989a. “Mechanics of overflow erosion on embank- ing of cohesive homogeneous earth dam with different cohesive
ments. I: Research activities.” J. Hydraul. Eng. 115 (8): 1040–1055. strength.” Sci. China Ser. E: Technol. Sci. 52 (10): 3024–3029. https://
https://doi.org/10.1061/(ASCE)0733-9429(1989)115:8(1040). doi.org/10.1007/s11431-009-0275-1.

© ASCE 04020014-16 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014


Zhao, G., P. J. Visser, and P. Peeters. 2013a. Large-scale embankment Zhao, G., P. J. Visser, R. Yankai, and W. S. J. Uijttewaal. 2015. “Flow
breach experiments in flume. Rep. of Delft Univ. of Technology. hydrodynamics in embankment breach.” J. Hydrodyn. Ser. B 27 (6):
Antwerp, Belgium: Flanders Hydraulics Research and Rijkswaterstaat. 835–844. https://doi.org/10.1016/S1001-6058(15)60546-7.
Zhao, G., P. J. Visser, P. Peeters, and J. K. Vrijling. 2013b. “Headcut Zhu, Y., P. J. Visser, J. K. Vrijling, and G. Wang. 2011. “Experimental
migration prediction of the cohesive embankment breach.” Eng. Geol. investigation on breaching of embankments” Sci. China Technol. Sci.
164 (Sep): 18–25. https://doi.org/10.1016/j.enggeo.2013.06.012. 54 (1): 148–155. https://doi.org/10.1007/s11431-010-4208-9.
Downloaded from ascelibrary.org by Lulea University of Technology on 01/29/20. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 04020014-17 J. Hydraul. Eng.

J. Hydraul. Eng., 2020, 146(4): 04020014

You might also like