Sei sulla pagina 1di 15

Engineering Geology 251 (2019) 48–62

Contents lists available at ScienceDirect

Engineering Geology
journal homepage: www.elsevier.com/locate/enggeo

Simulation of interactions between debris flow and check dams on three- T


dimensional terrain

Hong-Xin Chen, Jin Li, Shi-Jin Feng , Hong-Yu Gao, Dong-Mei Zhang
Key Laboratory of Geotechnical and Underground Engineering of Ministry of Education, Department of Geotechnical Engineering, Tongji University, Shanghai 200092,
China

A R T I C LE I N FO A B S T R A C T

Keywords: Debris flows are rapid gravity-driven flows of sediment-water mixture, which can be greatly destructive due to
Debris flow its huge volume, high velocity and large impact force. Check dam is essential protective structure for controlling
Check dam debris flow. However, it is a challenging work to assess the interactions between debris flow and check dam,
Destruction especially when involving complex topography and dam destruction. A numerical method was developed in this
Fluid-structure interaction
study to investigate the interactions between debris flow and check dam on three-dimensional terrain. The debris
Stava tailings dam
flow and check dam were simulated by Smoothed Particle Hydrodynamics (SPH) method and Finite Element
Method (FEM), respectively. The method was validated by a dam break problem and a granular flow flume test.
An actual debris flow event originated from failure of tailings dams on 19 July 1985 in Stava, Italy was simulated
as an example. The results indicate that the proposed method is a practical tool to simulate the runout char-
acteristics of debris flow (e.g., flow velocity, flow depth, impact area) and interactions between debris flow and
check dam (e.g., impact force, destruction of check dam, interception by check dam). Given similar total dam
volume, increasing the number of dams will improve the hazard mitigation effect. Moreover, it is recommended
to construct dams in downstream area with straight channel. This study will contribute to a better understanding
of the flow-structure interaction and is helpful for rational design of check dams.

1. Introduction flow and check dam is a challenging work, due to the complicated
runout characteristics of debris flow and the difficulty in accurately
Debris flows are rapid gravity-driven flows of sediment-water mix- describing the impact on check dam especially when involving complex
ture, which can be greatly destructive due to its huge volume, high topography and dam destruction.
velocity and large impact force, which can pose great danger to the The runout characteristics of debris flow have been comprehen-
downstream people and infrastructure. For example, a giant cata- sively investigated by field investigation (e.g., Tang et al., 2011; Xu
strophic debris flow attacked Zhouqu, China, destroying or dama- et al., 2012; Chen et al., 2012, 2014; Zhang et al., 2014; Fan et al.,
ging > 5500 buildings and causing 1765 fatalities (Tang et al., 2011). 2018), experimental test (e.g., Iverson et al., 2011; Zhou et al., 2013,
Similar tragedies caused by debris flow have been reported by nu- 2015), and numerical simulation (e.g., McDougall and Hungr, 2004;
merous researchers (e.g., Chandler and Tosatti, 1995; Revellino et al., Pastor et al., 2009; Luna et al., 2012; Chen and Zhang, 2015; Gao et al.,
2004; Xu et al., 2012; Zhang et al., 2014; Zhou et al., 2015; Ouyang 2016; Zhang and Matsushima, 2016; Chen et al., 2017; Ouyang et al.,
et al., 2017; Wei et al., 2018). 2017; Shen et al., 2017; Braun et al., 2018) in the past decades, mainly
In order to mitigate the hazard, protective structures, such as check focusing on the variation of flow depth, flow velocity, inundation area
dam (Shieh et al., 2007; Mizuyama, 2008; Liu et al., 2017), flexible and runout distance. These works provide basis for further studying the
barrier (Wendeler et al., 2007; Leonardi et al., 2016), deposition basin flow-structure interactions. Impact force of debris flow is an important
(Zollinger, 1985), are installed at different positions along the potential index for the design of protective structure. Several laboratory flume
flow path. Among those, check dam is the most widely used, which has tests have been carried to get the impact force of debris flow (Jiang and
to be rationally designed to resist the dynamic force caused by debris Towhata, 2013; Scheidl et al., 2013; Cui et al., 2015; Song et al., 2017;
flow. However, reliable evaluation of the interactions between debris Zhou et al., 2018). However, the tested barriers were rigid and the


Corresponding author.
E-mail addresses: chenhongxin@tongji.edu.cn (H.-X. Chen), 1150102@tongji.edu.cn (J. Li), fsjgly@tongji.edu.cn (S.-J. Feng),
1630533@tongji.edu.cn (H.-Y. Gao), dmzhang@tongji.edu.cn (D.-M. Zhang).

https://doi.org/10.1016/j.enggeo.2019.02.001
Received 15 August 2018; Received in revised form 26 January 2019; Accepted 2 February 2019
Available online 06 February 2019
0013-7952/ © 2019 Elsevier B.V. All rights reserved.
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

destruction process can be hardly observed; moreover, laboratory tests dviα


N
⎛ σi
αβ σ jαβ ⎞ ∂Wij F
normally simplify the topography, but the boundary effect and size = ∑ mj ⎜ + + i
dt ρi2 ρj2 ⎟ ∂x iβ mi (2)
effect may significantly influence the test results. So some field tests j=1 ⎝ ⎠
have been conducted to obtain valuable in-situ data, such as at Ill- where the subscript i represents the concerned particle and the sub-
graben torrent in Switzerland (Wendeler et al., 2007) and Jiangjia script j represents a neighbor particle in the influence domain; t is the
Ravine in China (Hu et al., 2011). But it is quite difficult and dangerous time; N is the total number of particles within the influence domain of
to get real-time record of impact force during the debris flow (Dai et al., the particle i; m and ρ are the mass and density of the particle, re-
2017). spectively; σ is the stress tensor of the particle; Wij = W(xi-xj, h) is the
Therefore, numerical simulation is adopted by some researchers to smoothing Kernel function and h is the smoothing length determining
study the flow-dam interactions. For example, Remaître et al. (2008) the influence domain of the smoothing function; vij is the relative ve-
adopted a one dimensional model to assess the influence of check dams locity vector between particles i and j; the superscripts α and β represent
on debris flow runout intensity; Kwan et al. (2015) conducted a staged the coordinate directions; F denotes the external forces such as gravity,
debris mobility analysis that accounts for the effects of multiple check basal friction, and interaction forces with check dam, which are simply
dams; Dai et al. (2017) developed a numerical model using the SPH applied to the particles without using any SPH approximation.
method to investigate the impact force of debris flow. Although the
above works can simulate the interactions between debris flow and
2.2. Flow resistance
check dam at catchment scale, the destruction process can not be
modeled since the structure is assumed to be unbreakable. In fact, check
The concerned resistance to debris flow movement includes two
dams can be seriously damaged or even destroyed by debris flow, which
parts in this study. The first one is basal friction at the interface between
has been widely observed in field (Remaître et al., 2008; Tang et al.,
debris flow and ground surface, which is characterized by coefficient of
2011; Xu et al., 2012; Chen et al., 2015). The destruction may in turn
friction between debris flow and ground and evaluated by Mohr-
substantially alter the runout characteristics of debris flow.
Coulomb model. The basal friction is one of the external forces in Eq.
The objective of this study is to investigate the interactions between
(2). The second one is the internal flow resistance in debris flow ma-
debris flow and check dam on three-dimensional terrain using numer-
terial, which is controlled by the adopted material model. There are
ical method, especially the destruction of check dam. The proposed
about 140 material models in LS-DYNA (Hallquist, 2006), and the fol-
method is firstly validated by a dam-break problem and a flume test. A
lowing two ones are suitable for simulating flow materials.
debris flow originated from tailings dam failure on 19 July 1985 in Italy
The first one is the null material (*MAT_NULL), which has no yield
was then simulated as a case study. Three scenarios were concerned,
stress and behaves in a fluid-like manner. In this model, the internal
namely, with no check dam, with rigid check dam(s), and with de-
flow resistance is characterized by viscous shear stress, which can be
structible check dam(s). This study will contribute to a better under-
defined as follows:
standing of the flow-structure interaction and is helpful for rational
design of check dams. σ αβ = 2με αβ
̇ (3)

where μ is the dynamic viscosity; ε ̇ is the deviatoric strain rate. The


2. Methodology viscous stress serves as one of the interaction forces between different
particles in Eq. (2), which resists the movement of particles.
LS-DYNA is a powerful numerical platform developed by Livermore The other material model is the elastic-plastic hydrodynamic ma-
Software Technology Corporation (Hallquist, 2006), which has been terial (*MAT_ELASTIC_PLASTIC_HYDRO, abbreviated as MEPH), which
successfully applied to simulate debris flow runout processes (e.g., represents a continuous medium having both fluid properties and
Konuk et al., 2006; Kwan et al., 2015; Koo, 2017; Koo et al., 2017), elastoplastic solid properties. The MEPH model is applicable for a
seismic response of structures (Ding et al., 2006; Lee and Chang, 2012) variety of materials, including those showing pressure dependent yield
and blasting-induced response of soil and rock (Ma and An, 2008; An behaviors (Hallquist, 2006), thus it can be used for representing debris
et al., 2011). Due to its capability of describing dynamic and high flow with high solid concentration. In this model, the internal flow
strain-rate problems, it was adopted to simulate the movement of debris resistance is characterized by yield stress σy. In a time step, a trial value
flow and destruction of check dams in this study. Considering the fea- of deviatoric stress is computed first:
tures of debris flow and check dam, the debris flow was modeled using
the Smoothed Particle Hydrodynamics (SPH) method, and the other snαβ+ 1 = snαβ + Rαβ + 2Gε αβ
̇ dt (4)
elements were modeled using the Finite Element Method (FEM). where the superscript ‘-’ denotes a trial value; the subscript n denotes
the time step; sn is the deviatoric stress at the n-th time step; R is the
2.1. SPH method rotation matrix containing the corotational basis vectors (Hallquist,
2006); G is the shear modulus.
SPH is a meshless method defined by a number of particles with a A trial value of effective stress is then calculated as
field around them. The main advantage of SPH method over classical 3 0.5
CFD methods is that a numerical grid is not needed, which helps to s = ⎛ snαβ+ 1 snαβ+ 1 ⎞
⎝2 ⎠ (5)
avoid the limitations of mesh tangling encountered in extreme de-
formation scenarios especially when coupled with the finite element If s is smaller than the yield stress, the material falls into elastic state
method. and the plastic strain increment is zero. If s exceeds the yield stress,
In SPH method framework, the computational domain is discretized plastic strain increment occurs, and the yield stress is updated as
into numerous particles on which the governing equations are solved. (σy )n + 1 = (σy )n + Eh Δε p (6)
In this study, Navier-Stokes equations were adopted as the governing
equations for debris flow movement, which can be written as a set of where Eh is the plastic hardening modulus; Δεp is the plastic strain in-
equations in the form of particle approximation as follows: crement. The trial stress will be scaled back to the yield surface, and the
updated deviatoric stress is
N
dρi ∂Wij
= ∑ mj vijβ (σy )n + 1
dt j=1 ∂x iβ (1) snαβ+ 1 = snαβ+ 1
s (7)

49
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

The stress tensor of the particle in Eq. (2) is then updated as direction, the automatic contact module is adopted. Apart from the
normal contact force, frictional force also exists at the interface, which
σnαβ+ 1 = snαβ+ 1 − pn + 1 δ αβ (8) is assessed based on the coefficient of friction. On one hand, the in-
where p is the pressure; δ is the Kronecker delta. teraction forces may deform the check dam; on the other hand, the
It should be noted that an equation of state (EOS) should be adopted interaction forces will serve as external forces in Eq. (2), changing the
to compute the pressure term in both material models. The widely used movement of debris flow.
Gruneisen EOS is chosen in this study:
γ0
2.4. Structure destruction

p=
ρ0 C 2θ ⎡1 + 1 −
⎣ ( 2 )θ − a 2
θ⎤
2 ⎦
+ (γ0 + aθ) E
θ2 θ3
2 The destruction of check dam is considered through element erosion
⎡1 − (S1 − 1) θ − S2 θ + 1 − S3 (θ + 1)2 ⎤
⎣ ⎦ (9) algorithm. An illustration of the destruction is shown in Fig. 1b. Debris
flow impacts the check dam, and stress state changes in elements of
p = ρ0 C 2θ + (γ0 + aθ) E (10) check dam. Once the changes develop to a certain extent, element
where p is the pressure; ρ0 is the initial density; C is the intercept of the erosion occurs. Several failure criteria are available in the platform,
curve corresponding to the adiabatic speed of sound; S1, S2 and S3 are including critical pressure, critical principal stress, critical principal
the fitting coefficients; γ0 is the Gruneisen coefficient; a is the volume strain and so on, one or more of them could be selected as needed and
correction coefficient to γ0; θ = 1/V–1, where V is the relative volume, they work independently (Hallquist, 2006). In this study, critical prin-
θ > 0 for compressed state, θ < 0 for expanded state; E is the internal cipal strain was adopted:
energy per initial volume. Eq. (9) is for compressed state, and Eq. (10) is ε1 ≥ εf (13)
for expanded state. It allows a precise propagation of the pressure wave
in a computationally efficient manner. where ε1 is the maximum principal strain; εf is the principal strain at
failure.
2.3. Fluid-structure interactions If Eq. (13) is satisfied, the element would be deleted from calcula-
tion. The deletion process is irreversible, which means that the deleted
When debris flow contacts check dam, interactions occur between element would not be involved in the later calculation while the other
them (Fig. 1). In order to simulate the interactions, coupling between normal elements still work. This makes contact between SPH particles
the FEM elements (check dam) and the SPH particles (debris flow) is and check dam elements continue. So the check dam would be damaged
realized by using contact algorithm. The penalty method is chosen as if some elements are removed or even destroyed if most elements are
the contact algorithm to evaluate the interactions. In this method, removed. Therefore, the method is applicable for simulating the de-
debris flow consisting of a set of SPH particles is designated as the slave struction process of check dam. Although the destruction process is
nodes, and the surface of check dam is designated as the master surface simplified and failure elements are deleted instead of being cracked or
(Fig. 1a). At each time step, the slave nodes are checked to see whether separated, it can also reflect structure destruction to a certain extent.
they penetrate into the master surface. If penetration occurs, an inter-
face contact force is introduced: 2.5. Solution procedures
F = kl (11)
An illustration of the calculation program in this paper is shown in
where k is the contact stiffness and l is the penetration depth. This is Fig. 2. In a debris flow-check dam interaction problem, the ground and
physically equivalent to placing an interface spring between the pene- check dams are built using FEM, and debris flow mixture is simulated
trating nodes and the contact surface (see Fig. 1a). The contact stiffness using SPH particles which are assigned with initial properties such as
is calculated as density, location and velocity.
αKA2 m ⎫ In each time step, a neighbor list for each particle is created first
k = max ⎧ , SOFSCL based on the smoothing length; then the neighbor particles are sorted

⎩ V 2∆t 2 ⎬
⎭ (12)
and the density, strain rate, pressure and stress of each particle are
where α is a scale factor; K is the bulk modulus of check dam, which is calculated based on Eq. (1) and material model; afterwards, the particle
calculated based on the material properties; A is the area of master force is evaluated. When the SPH particles touch the check dam ele-
surface; V is the volume of check dam element; SOFSCL is the scale ments as shown in Fig. 1a, interactions occur between them following
factor for the Soft Constraint Penalty Formulation; m is the nodal mass the principle introduced in Section 2.3. For the SPH particles, the ac-
and Δt is the time step. Since it is difficult to determine the contact celeration can be calculated based on the particle force and external

Fig. 1. Illustration of interactions between debris flow and check dam.

50
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

Fig. 2. Calculation program of the present method.

(a)

(b)
10

8
Water depth (m)

4 Point 1 in this paper


Gabutti Point 1
MacCormack Point 1
2 Point 2 in this paper
Gabutti Point 2
MacCormack Point 2
0
0 1 2 3 4 5 6 7
Time (s)
Fig. 3. (a) Water depth at 7.5 s after the dam breached computed by the present method; (b) comparison of the computed water depth between the present method
and two numerical schemes reported by Fennema and Chaudhry (1990).

forces based on Eq. (2), and the velocity and position can be then up- adopted for time integration to ensure numerical stability.
dated. For the check dam elements, given the interaction forces, the
dynamic equation is formulated. On the basis of the constructed stiff-
ness matrix and mass matrix, the displacement can be solved, and the 3. Model validation
strain and stress can be then obtained. If the element does not fail, it
remains working. Otherwise, the element is deleted, and the stiffness 3.1. Test 1: dam break
matrix and mass matrix will be changed correspondingly in the next
time step. The classical Courant-Friedrich-Lewy (CFL) condition is A dam break problem reported by Fennema and Chaudhry (1990)
was adopted here to test the performance of the method in simulating

51
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

Table 1 reservoir water was 10 m, and that of the tail water was 5 m. The re-
Adopted parameters for numerical simulation. servoir water was initially retained by a dam. At t = 0, the dam was
Parameter Meaning Value Reference assumed to fail instantaneously and the breach width was 75 m. The
water was modeled using MAT_NULL, which is perfect to model water-
Parameters of Gruneisen EOS like fluid, with a density of 1000 kg/m3 and a dynamic viscosity of
C (m/s) Sound speed 1480 Liu et al. (2003)
0.001 Pa·s. The initial spacing of SPH particles was 1.25 m. The adopted
S1 Fitting coefficient 2.56 Liu et al. (2003)
S2 Fitting coefficient 1.986 Liu et al. (2003)
parameters for Gruneisen EOS are summarized in Table 1. The com-
S3 Fitting coefficient 1.2268 Liu et al. (2003) puted surface profile of water at 7.5 s after the dam breached is shown
γ0 Gruneisen coefficient 0.5 Liu et al. (2003) in Fig. 3a, which was almost the same as that reported by Fennema and
a Volume correction coefficient 0 Liu et al. (2003) Chaudhry (1990). Two points (Points 1 and 2 in Fig. 3a) were selected
Parameters of contact to further investigate the variation of water depth. The comparison
μ1 Coefficient of friction between debris 0.045 between the present method and two numerical schemes reported by
flow and ground
Fennema and Chaudhry (1990) is shown in Fig. 3b, there was slight
μ2 Coefficient of friction between debris 0.045
flow and check dam difference at the beginning but the results agreed reasonably well later.
α Scale factor for the interface stiffness 0.1 Hallquist (2006) The reason for the difference may be that Fennema and Chaudhry
SOFSCL Scale factor for the Soft Constraint 0.1 Hallquist (2007) (1990) adopted depth-averaged shallow water equation and finite dif-
Penalty Formulation ference method, while this study conducted three-dimensional simula-
tion using SPH method.
three-dimensional flow, which is illustrated in Fig. 3a. The computa-
tional domain was a channel with a length of 200 m and a width of
200 m. The boundary was assumed to be frictionless. The depth of the

(a)

(b)
500
F1 (This paper)
F2 (This paper)
F3 (This paper)
400
F1 (Jiang and Towhata, 2013)
Impact force (N/m)

F2 (Jiang and Towhata, 2013)


F3 (Jiang and Towhata, 2013)
300

200

100

0
0.0 0.5 1.0 1.5 2.0
Time (s)
Fig. 4. (a) Schematic diagram of the flume test; (b) comparison of impact forces on the rigid wall.

52
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

Fig. 5. Comparison of flow paths and final deposition zones between the field investigation and numerical simulation, and locations of hypothetical check dams.

Fig. 6. Runout process of the debris flow: (a) time = 30 s; (b) time = 100 s; (c) time = 300 s; (d) time = 500 s.

3.2. Test 2: flow impact on rigid wall MAT_NULL. Since it was a dry flow, the dynamic viscosity was assumed
to be zero. The friction angles between the particles and the base, the
The granular flow flume test conducted by Jiang and Towhata retaining wall and the side wall were 25°, 21° and 15°, respectively
(2013) was adopted to validate the performance of the present method (Jiang and Towhata, 2013), which were adopted in the simulation. The
in simulating the interaction between debris flow and barrier. The side comparison of impact forces is shown in Fig. 4b. The results overall
view of the test apparatus is shown in Fig. 4a. The length, the width and showed good agreement between the simulation results and the test
the height of the flume were 2.63, 0.3 and 0.35 m, respectively. A rigid data.
wall perpendicular to the flume base was installed at the end of the
flume, and the impact force was measured by load cells on the wall. A
sliding mass composed of limestone gravel was initially retained at the
entrance of the flume, with a density of 1378 kg/m3. The initial spacing
of SPH particles was 0.02 m. The granular flow was also modeled with

53
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

35 4. Simulation of the debris flow originated from the failure of


This paper
Stava tailings dams on 19 July 1985
Takahashi (2007)
30
4.1. The debris flow event
Flow velocity (m/s)

25
A destructive debris flow was triggered by the collapse of tailings
20
dams in Stava (northern Italy) on 19 July 1985, which has been con-
cretely investigated in the past decades through field survey, inter-
15
pretation of aerial images, and numerical simulation (e.g., Muramoto
et al., 1986; Chandler and Tosatti, 1995; Luino and De Graff, 2012;
10
Pirulli et al., 2017). Two fluorite tailings dams were built on a slope, at
5 elevation ranging from 1330 to 1380 m. The upper dam collapsed first
and then the lower dam subsequently fell down. The tailings rushed out
0 immediately, then flowed downstream along the Stava Valley and fi-
0 500 1000 1500 2000 2500 3000 3500 nally deposited at the Avisio River (Fig. 5), traveling over 4.2 km. The
Distance (m) runout path and impact area of the debris flow were determined
through field survey and interpretation of aerial images. According to
Fig. 7. Comparison of flow velocity profiles. the survey after this event, 185,000 m3 sandy tailings and silty tailings
with a bulk density of 1900 kg/m3 (Chandler and Tosatti, 1995) and a
Table 2 volumetric sediment concentration of 0.476 (Takahashi, 2007) were
Summary of interactions between debris flow and check dam(s) without con- released. Most houses in Stava Valley and 47 buildings along the stream
sidering dam destruction. in Tesero were destroyed and 268 people were killed (Muramoto et al.,
Case Location H (m) Vd (m3) Fmax (N) λm Qmax T (s) 1986).
(m3/s) The event was simulated in this study since the information, espe-
cially the digital terrain, is public and detailed, making it possible for
Base case None None None None None 969 162
other interested researchers to follow up. Although no check dam ex-
1 1 10 2500 4.11 × 108 23.8% 937 206
2 1 15 4500 5.24 × 108 46.3% 890 222 isted in the event, some hypothetical check dams were adopted to in-
3 1 20 7000 6.02 × 108 78.5% 609 243 vestigate the runout characteristics of debris flow and the flow-dam
4 5 10 1500 1.29 × 108 46.8% 672 206 interactions. Close check dam, which is one of the most widely used
5 5 15 3000 1.66 × 108 76.5% 406 228 check dams, was adopted in this study. Totally three scenarios were
6 9 10 2000 7.78 × 107 78.6% 422 253
7 9 15 3750 9.62 × 107 92.4% 375 291
simulated, including the scenario with no check dam (actual scenario),
8 1, 2 10 5000 1–4.20 × 108 74.6% 469 247 scenario with the assumed rigid check dam(s) (not destructible) and
2–1.38 × 108 scenario with the assumed concrete check dam(s) (destructible).
9 1, 2, 3 7.5 5062.5 1–3.50 × 108 96.5% 102 273
2–1.40 × 108
4.2. Runout characteristics of the debris flow with no check dam
3–7.07 × 107
10 4, 5 10 3000 4–1.76 × 108 85.1% 219 285
5–6.08 × 107 Digital elevation data with a spatial resolution of 1 m was firstly
11 4, 5, 6 7.5 3187.5 4–1.53 × 108 96.1% 37 425 downloaded from the GIS web of Trentino Geocartographic Portal of
5–5.56 × 107 the Autonomous Province of Trento. Considering the computational
6–3.70 × 107
12 8, 9 10 4000 8–9.75 × 107 100% – –
domain is large, a too small cell size will lead to enormous computa-
9–2.23 × 107 tional cost, so a digital elevation file with a spatial resolution of 10 m
13 7, 8, 9 7.5 3937.5 7–7.33 × 107 100% – – was obtained from the original data in ArcGIS.
8–5.54 × 107 The Stava Valley did not suffer from significant erosion (Pirulli
9–9.67 × 106
et al., 2017), so erosion process was not considered in this study and the
14 1, 5, 9 7.5 3937.5 1–3.46 × 108 97.3% 4 508
5–7.42 × 107 ground was modeled with rigid shell elements. The rushing out tailings
9–3.05 × 107 were reconstructed by 11,119 SPH particles with an initial particle
spacing of 2.5 m. The tailings were assumed to be released instantly.
Note: H is the height of check dam; Vd is the volume of check dam(s); Fmax is the The mixture of silt, sand and water was modeled by the MEPH model
maximum impact force on check dam; λm is the mitigation ratio; Qmax is the with an average density of 1900 kg/m3. Major and Pierson (1992)
maximum discharge at the observation point; T is the arrival time of debris flow
presented a series of experimental data of yield stress for fine-grained
at the observation point.
slurries, which were similar to the debris flow material in this event. So
the data was adopted to estimate the yield stress of debris flow in this
6.0x108
Location 1, H = 10 m (Case 1) study. Considering Cv was 0.476 in this event, the yield stress was de-
Location 1, H = 15 m (Case 2) termined as 40 Pa. The shear modulus of debris flow was taken as
Location 5, H = 10 m (Case 4) 6 MPa, which is the typical value for debris flow reported by Koo
Location 5, H = 15 m (Case 5)
Impact force (N)

4.0x108 (2017). In fact, we found that the simulation results were not sensitive
Location 9, H = 10 m (Case 6)
Location 9, H = 15 m (Case 7) to the shear modulus, which was also reported by Koo (2017). The
plastic hardening modulus was taken as zero for simplicity, otherwise
2.0x108 the yield stress would increase to a greatly high level since the de-
formation during the debris flow movement was extremely large. As
mentioned in Section 2.2, coefficient of friction is another important
0.0
parameter for evaluating flow resistance. Since there was no available
0 50 100 150 200 250 300 data for the contact friction between the debris flow and the ground
Time (s) (μ1) in this event, the coefficient of friction was determined as 0.045
based on back analysis. The relatively low value of μ1 is mainly at-
Fig. 8. Evolution of impact force on single check dam at different locations.
tributed to the high fine content of the debris flow material. The low

54
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

Fig. 9. Interaction between debris flow and check dam of Case 2: (a) time = 52 s; (b) time = 55 s; (c) time = 59 s.

permeability may cause large excess pore water pressure at the base, estimate debris flow velocity in field survey (Hungr et al., 1984;
which can significantly reduce the flow resistance. The adopted para- Hürlimann et al., 2003; Chen et al., 2014). In this study, the velocity of
meters of Gruneisen EOS and the other contact parameters are sum- SPH particles was monitored and output directly and the average flow
marized in Table 1. velocity of debris flow front was calculated. The comparison of flow
The runout process of debris flow was simulated, as shown in Fig. 6. velocity profiles between the simulation result and the data reported by
The tailings dam collapsed, rushed out, and then flowed downward to Takahashi (2007) is shown in Fig. 7, which agreed reasonably well,
the Stava Valley in a short period (30 s). The flow velocity has reached a except for the beginning part. There are two reasons for the difference.
very high level, making it greatly destructive. Afterwards, the mass One reason is that the tailings were released instantaneously in the si-
flowed along the channel (Fig. 6b), forming a long strip of flow with a mulation, which was slightly different from the real scenario as afore-
high-speed front and a low-speed tail. The debris flow moved with a mentioned. Takahashi (2007) estimated the duration of the failure
decreasing speed and finally deposited at the confluence area between process by applying the Ritter's dam collapse function and reported that
the Stava Valley and the Avisio River (Fig. 6d). The comparison of flow the total volume of debris flow was released for about 13.2 s. Another
paths and final deposition zones between the field data (Muramoto reason is that there were numerous trees near the tailings dam. Luino
et al., 1986) and the simulation result is shown in Fig. 5. Good agree- and De Graff (2012) reported that hundreds of tall trees (spruce and
ment can be observed. larch) were literally cut down just above the roots, which would greatly
Apart from the impact area, flow velocity is also an important index reduce the debris flow velocity. But this factor was difficult to be
for evaluating the runout characteristics of debris flow. Two months considered in the simulation. The above results reveal that the present
after the event, Takahashi and his colleagues did the field survey. They method can well simulate debris flow at catchment scale and the
found obvious difference in flood marks between one on the left bank adopted parameters are reasonable.
and the other on the right bank, the level difference is called super-
elevation. This is the effect of centrifugal force. The cross-sectional 4.3. Runout characteristics of the debris flow considering the influence of
average velocity of debris flow was estimated using the phenomenon of check dam
super-elevation and the detailed information is enclosed in Takahashi
(2007). In fact, the super-elevation method has been widely adopted to In this part, the influence of check dam was investigated and rigid

55
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

Fig. 10. Interaction between debris flow and check dam of Case 7: (a) time = 125 s; (b) time = 127 s; (c) time = 131 s; (d) time = 135 s.

check dams were adopted. Hypothetical check dam(s) were set along noted that the dam height decreased with increasing dam number at the
the flow path before the debris flow occurred. In order to evaluate the same area in order to make the different cases have similar dam scale,
mitigation effect of check dams considering various heights, numbers which is beneficial for comparison. For simplicity, the thickness of all
and locations, totally 14 cases were simulated as summarized in the dams was set as 5 m. In the downstream area, an observation point
Table 2, and the locations of the check dams are shown in Fig. 5. Nine was selected to assess the arrival time and flow discharge of debris flow
possible locations were selected, which were distributed almost evenly. (Fig. 5).
Cases 1–7 considered single check dam set at different locations (up- The simulation results are also summarized in Table 2, including the
stream, midstream, downstream) with various heights (10, 15, 20 m). maximum impact force on check dam, the mitigation ratio of check
Cases 8–14 considered multiple (two or three) check dams. It should be dam, the maximum flow discharge and arrival time of debris flow at the

56
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

40 reached the check dam and interacted intensely (Fig. 9a). After inter-
Case 2 acting with check dam, the velocity of the upper flow part was still
Check dam at Location 1
Case 5 large. Then the debris flow ran up significantly (Fig. 9b), some of the
Case 7 material flew over the dam and kept traveling, some was retained by
30 Check dam at Location 5
the dam and turned back (Fig. 9c). The huge impact force was domi-
Flow velocity (m/s)

nated by the dynamic pressure. For Case 7, the velocity of debris flow
Check dam at Location 9 H = 15 m was not so large. The debris flow had slowed down before reaching the
20
dam (Fig. 10a), and the run-up phenomenon was not evident (Fig. 10b).
85 s Then the material stopped behind the check dam (Fig. 10c) and gra-
dually filled up the reservoir (Fig. 10d), thus the impact force gradually
increased (see Fig. 8) and the stable value was attributed to the hy-
10 drostatic pressure.
Fig. 11 shows the evolution of debris flow velocity considering
different dam locations. The flow velocity dropped suddenly when
debris flow reached check dam, and then rose gradually after a certain
0 period of time. It took more time for debris flow moving over down-
0 100 200 300 400
stream check dam, leading to larger arrival time (see Table 2). For
Time (s) example, the flow mass spent about 85 s to recover the flow velocity at
Location 9 (Case 7). As a result, the arrival time in Case 7 was up to
Fig. 11. Variation of frontal velocity of debris flow with time for difference
cases.
291 s, and the observed maximum flow discharge was 375 m3/s, which
was monitored according to the number of passing particles per unit
time. Compared with the base case, the affected people would have
observation point. The results indicate that the impact force increased about two more minutes to flee away and the intensity of debris flow
with increasing dam height but decreased with increasing distance would be reduced by approximately 60% in terms of flow discharge.
between the source area and the check dam. Fig. 8 shows the variation Mitigation ratio, which is defined as the ratio of the sediment vo-
of impact force on single check dam with time at different locations. At lume retained by check dam to the total debris flow volume, is the most
Location 1, the impact force on the check dam with a height of 15 m direct parameter reflecting the control function of check dam(s). It can
(Case 2) reached the maximum value of 5.24 × 108 N at about 50 s and be expected that the mitigation ratio increased with increasing dam
then decreased rapidly, followed by a smaller second peak before height as shown in Table 2 since the destruction of dam was not con-
reaching a stable value of about 1.0 × 108 N. A similar curve appeared sidered. In addition, the mitigation ratio increased with the increase of
for check dam at the same location with a height of 10 m (Case 1), distance between the source area and the check dam. As shown in
which only had a maximum value of 4.11 × 108 N. The peaks of the Table 2, the debris flow intensity showed a decreasing trend with in-
curves for Location 5 (Cases 4 and 5) were much smaller but no second creasing dam number. For example, the mitigation ratios of Case 5,
peak appeared. It is noteworthy that the eventual stable values for Case 10 and Case 11 were 76.5%, 85.1% and 96.1%, respectively, while
Locations 1 and 5 with the same dam height were almost the same. As the dam scales were similar (about 3000 m3). Similar rules were also
for Location 9 (Cases 6 and 7), the impact force increased gradually found at the other locations. Therefore, multiple barriers were more
with time and no obvious peak appeared. effective because the energy gain of debris flow during overflows was
The detailed impact process is shown in Figs. 9 and 10 to further limited (Kwan et al., 2015). In particular, the mitigation ratios in Cases
investigate the interactions. For Case 2, the high-speed debris flow 12 and 13 were up to 100%, which means the debris flow was totally

Fig. 12. Comparison of final deposition zones among typical cases: (a) base case with no check dam; (b) Case 1 with one upstream check dam; (c) Case 13 with three
downstream check dams.

57
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

Fig. 13. Influence of super-elevation: (a) check dam constructed at Location 1; (b) check dam constructed at 200 m upstream from Location 1.

Table 3 controlled by the check dams. Fig. 12 shows a comparison of the final
Summary of interactions between debris flow and check dam(s) considering deposition zones among three typical cases. The debris flow was com-
dam destruction. pleted intercepted by the three dams (Case 13), forming a large sedi-
Case Location H (m) εf λR (%) λm (%) Qmax (m3/s) T (s) mentary area at the Stava Valley, while much debris flow reached the
Avisio River when there was no check dam (base case) or one upstream
15 9 15 0.1 35.1 0 750 195 check dam (Case 1).
16 9 15 0.2 18.8 3.0 656 239
In addition, debris flow sometimes shows a super-elevation phe-
17 9 15 0.3 2.9 0 813 269
18 1 15 0.2 6.5 1.2 984 199 nomenon when it goes through the curved channel at high speed. It is
19 5 15 0.2 40 31.9 688 210 relatively difficult to predict the debris flow path at the bend. In the
20 1, 2, 3 7.5 0.2 1–10 0 1094 198 Stava event, debris flow drastically changed the direction before en-
2–0 tering the Stava Valley. A check dam was assumed to be constructed at
3–6.7
21 4, 5, 6 7.5 0.2 4–7.6 47.8 500 235
the turning point, about 200 m upstream from Location 1 in order to
5–0.8 investigate the influence of super-elevation. As shown in Fig. 13a, the
6–100 check dam can effectively intercept the debris flow if it was located at
22 7, 8, 9 7.5 0.2 7–1.5 99.4 5 304 Location 1; however, in Fig. 13b considerable debris flow easily flowed
8–100
over the check dam on the right hand side due to super-elevation. So it
9–100
23 1, 5, 9 7.5 0.2 1–4.8 52.5 359 334 is suggested to construct check dams at straight channel for avoiding
5–3.8 super-elevation.
9–44.9

Note: εf is the principal strain of check dam material at failure; λR is the ratio of 4.4. Destruction of check dam
remaining check dam.

In order to investigate the influence of dam destruction, the check


dams were assumed to be made of concrete, which was modeled by

Fig. 14. Moment of check dam being destroyed considering different dam strengths: (a) εf = 0.1 in Case 15; (b) εf = 0.2 in Case 16; (c) εf = 0.3 in Case 17.

58
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

Fig. 15. Destruction process of check dam in Case 18: (a) time = 50 s; (b) time = 51 s; (c) time = 52 s; (d) time = 53 s.

Fig. 16. Residual structure of check dams in different cases: (a) Case 20; (b) Case 21; (c) Case 22.

MAT_CSCM_CONCRETE (Continuous Surface Cap Model, which is ab- ratios were all close to zero in these three cases. In Cases 18–23, εf was
breviated as CSCM here). The model is appropriate for predicting the 0.2. The destruction process of Case 18 was taken as an example to
dynamic performance of concrete used in safety structures (Murray further understand the failure mechanism, as shown in Fig. 15. In
et al., 2007). The density and unconfined compressive strength of the Fig. 15a, the debris flow material has interacted with the check dam,
concrete were 2500 kg/m3 and 38.5 MPa, respectively, which are the mainly in the lower part where the principal strain had reached very
design parameters of C60 concrete. In this part, totally 9 cases were high level. In Fig. 15b, the principal strain further increased. In Fig. 15c,
simulated and are summarized in Table 3. some elements had reached the critical state and were deleted from the
The influence of dam strength in terms of different εf was considered simulation. Finally, the whole dam was totally destroyed as shown in
in Cases 15–17. Considering the influence of steel and the purpose of Fig. 15d. Comparing Cases 16, 18 and 19, downstream check dam also
parametric study, a larger range of εf (0.1–0.3) was adopted in this led to larger arrival time, which was the same as the scenario without
study compared with the value reported by Murray et al. (2007). With considering dam destruction.
the increase of dam strength, the ratio of remaining check dam de- Situations with multiple check dams were more complicated. When
creased but the arrival time increased. The moment of check dam being the check dams were constructed at upstream area (Case 20 in
destroyed in these cases is shown in Fig. 14. In Case 15, the check dam Fig. 16a), the three check dams were completely destroyed. However,
was destroyed instantaneously due to its low strength. The debris flow the dam at Location 6 in Case 21 (Fig. 16b) as well as those at Locations
flowed through the break quickly and 35.1% of the dam remained 8 and 9 in Case 22 (Fig. 16c) still partially or completely remained after
(Fig. 14a). The high-strength dam blocked debris flow for a longer being impacted. The mitigation ratio in Case 22 reached 99.4%, which
period, but the dam collapsed as a whole because of the higher integrity achieved excellent performance. Therefore, even if the destruction of
(Case 17 in Fig. 14c). The medium strength case (Case 16) was in be- check dam was considered, multiple check dams were favorable for
tween Case 15 and Case 17. Due to the dam destruction, the mitigation hazard mitigation and it is recommended to construct the dams in

59
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

8
(a) Location 1 - Rigid (b) 1.6x10 Location 1 - CSCM
3.0x108 Location 2 - Rigid Location 2 - CSCM
Location 3 - Rigid Location 3 - CSCM
1.2x108
Impact force (N)

Impact force (N)


Case 9 Case 20
2.0x108
8.0x107

1.0x108
4.0x107

0.0 0.0
40 60 80 100 120 40 60 80 100 120
Time (s) Time (s)
8 8
(c)1.6x10 Location 4 - Rigid (d)1.2x10 Location 4 - CSCM
Location 5 - Rigid Location 5 - CSCM
Location 6 - Rigid Location 6 - CSCM
1.2x108 9.0x107
Impact force (N)

Impact force (N)


Case 11 Case 21

8.0x107 6.0x107

4.0x107 3.0x107

0.0 0.0
70 105 140 175 70 105 140 175
Time (s) Time (s)
8 8
(e)1.0x10 Location 7 - Rigid
(f) 1.0x10 Location 7 - CSCM
Location 8 - Rigid Location 8 - CSCM
Location 9 - Rigid Location 9 - CSCM
7.5x107 7.5x107
Impact force (N)

Impact force (N)

Case 13 Case 22

5.0x107 5.0x107

2.5x107 2.5x107

0.0 0.0
100 150 200 250 300 100 125 150 175 200
Time (s) Time (s)

Fig. 17. Variation of impact force with time considering multiple check dams: (a) upstream dams with rigid material (Case 9); (b) upstream dams considering
destruction (Case 20); (c) midstream dams with rigid material (Case 11); (d) midstream dams considering destruction (Case 21); (e) downstream dams with rigid
material (Case 13); (f) downstream dams considering destruction (Case 22).

downstream area. Table 3). Therefore, the distance between check dams should not be too
The impact forces in this section were compared with those in cases long. It is obvious that the check dam design in Case 22 was still the
without considering destruction (Fig. 17). It is obvious that in the same most effective in Table 3. The debris flow was almost completely con-
region, the impact force on the first two check dams (Locations 1, 2, 4, trolled by the check dams in this case, even though dam destruction was
5, 7, 8) considering dam destruction was smaller than that without considered.
considering dam destruction except Location 8. The rule for the last
check dam (Locations 3, 6, 9) was contrary. The reason is as follows. If 5. Limitations of the method
the first two dams were rigid, they sustained most impact energy. On
the contrary, the first two dams were damaged or even destroyed if The method is able to simulate interactions between debris flow and
destructible dam material was used, it can be expected that the last dam check dam considering the destruction process. Although some pre-
would bear more impact energy compared with the scenario without liminary results were obtained, the method has the following limita-
considering dam destruction. Moreover, for rigid check dams, the im- tions due to the simplifying assumptions.
pact force exerted on upstream dam was generally larger than that on
downstream one (Fig. 17a, c, e), which was not valid for destructible (1) The debris flow mixture was simplified as a single-phase fluid and
check dams (Fig. 17b, d, f). modeled by MEPH, which can not well account for the viscous ef-
In the above analysis, the check dams were located at upstream or fect. More suitable material models should be implemented through
midstream or downstream area. In Case 14 and Case 23, three check secondary development.
dams were constructed at upstream area (Location 1), midstream area (2) The destructible check dams were modeled by concrete material,
(Location 5), downstream area (Location 9), respectively. The control and the dam configuration was relatively simple. Future work is
function was still acceptable if destruction was ignored (Case 14). But if needed to study the more detailed destruction process of dam by
the destruction was considered (Case 23), the first two check dams were adopting more realistic check dam, which will contribute to a better
severely damaged and the mitigation ratio was only 52.5% (see understanding of the failure mechanism.

60
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

(3) In this study, the simulation was only based on the case of Stava Cui, P., Zeng, C., Lei, Y., 2015. Experimental analysis on the impact force of viscous debris
tailings dam failure. The method should be applied to more ex- flow. Earth Surf. Process. Landf. 40 (12), 1644–1655.
Dai, Z., Huang, Y., Cheng, H., Xu, Q., 2017. SPH model for fluid-structure interaction and
amples in the future to investigate the influence of topography, its application to debris flow impact estimation. Landslides 14 (3), 917–928.
which will be substantially helpful to the rational planning of check Ding, J.H., Jin, X.L., Guo, Y.Z., Li, G.G., 2006. Numerical simulation for large-scale
dams in engineering practice. seismic response analysis of immersed tunnel. Eng. Struct. 28 (10), 1367–1377.
Fan, R.L., Zhang, L.M., Wang, H.J., Fan, X.M., 2018. Evolution of debris flow activities in
Gaojiagou Ravine during 2008–2016 after the Wenchuan earthquake. Eng. Geol. 235,
6. Summary and conclusions 1–10.
Fennema, R.J., Chaudhry, M.H., 1990. Explicit methods for 2-D transient free surface
flows. J. Hydraul. Eng. 116 (8), 1013–1034.
A numerical method was developed to investigate the interactions Gao, L., Zhang, L.M., Chen, H.X., Shen, P., 2016. Simulating debris flow mobility in urban
between debris flow and check dam in this paper. The debris flow and settings. Eng. Geol. 214, 67–78.
check dam were simulated by SPH method and FEM, respectively. The Hallquist, J.O., 2006. LS-DYNA Theory Manual. Livermore Software Technology
Corporation, USA.
method was effectively validated by a dam break problem and a
Hallquist, J.O., 2007. LS-DYNA Keyword User's Manual. Livermore Software Technology
granular flow flume test. An actual debris flow event originated from Corporation, USA.
failure of tailings dams on 19 July 1985 in Stava, Italy was simulated as Hu, K., Wei, F., Li, Y., 2011. Real-time measurement and preliminary analysis of debris-
an example, three different scenarios were considered. Some major flow impact force at Jiangjia Ravine, China. Earth Surf. Process. Landf. 36 (9),
1268–1278.
conclusions are drawn as follows. Hungr, O., Morgan, G.C., Kellerhals, R., 1984. Quantitative analysis of debris torrent
hazards for design of remedial measures. Can. Geotech. J. 21 (4), 663–677.
(1) The proposed method is a practical tool to investigate fluid-struc- Hürlimann, M., Rickenmann, D., Graf, C., 2003. Field and monitoring data of debris-flow
events in the Swiss Alps. Can. Geotech. J. 40 (1), 161–175.
ture interaction. The runout characteristics of debris flow (e.g., flow Iverson, R.M., Reid, M.E., Logan, M., LaHusen, R.G., Godt, J.W., Griswold, J.P., 2011.
velocity, flow depth, impact area) and interactions between debris Positive feedback and momentum growth during debris-flow entrainment of wet bed
flow and check dam (e.g., impact force, destruction of check dam, sediment. Nat. Geosci. 4 (2), 116–121.
Jiang, Y.J., Towhata, I., 2013. Experimental study of dry granular flow and impact be-
interception by check dam) can be well simulated. havior against a rigid retaining wall. Rock Mech. Rock. Eng. 46 (4), 713–729.
(2) In upstream area, the debris flow generally moved at high speed, Konuk, I., Yu, S., Evgin, E., 2006. Application of the ALE FE method to debris flows. WIT
the flow-dam interaction was quite intense. The peak impact force Trans. Ecol. Environ. 90, 47–57.
Koo, R.C.H., 2017. 3D debris mobility assessment using LS-DYNA. In: GEO Report No.
on check dam was obvious and the huge impact force was domi- 325. Geotechnical Engineering Office, Hong Kong.
nated by dynamic pressure. On the contrary, in downstream area, Koo, R.C.H., Kwan, J.S., Lam, C., Goodwin, G.R., Choi, C.E., Ng, C.W.W., Pun, W.K., 2017.
debris flow moved at relatively small speed, the impact was much Back-analysis of geophysical flows using three-dimensional runout model. Can.
Geotech. J. 55 (8), 1081–1094.
weaker and the reservoir behind dam was filled gradually if the
Kwan, J.S.H., Koo, R.C.H., Ng, C.W.W., 2015. Landslide mobility analysis for design of
dam was not destroyed. multiple debris-resisting barriers. Can. Geotech. J. 52 (9), 1345–1359.
(3) Debris flow hazard was effectively controlled by check dams, which Lee, K.Z.Z., Chang, N.Y., 2012. Predictive modeling on seismic performances of geosyn-
significantly reduced the flow velocity, intercepted debris flow and thetic-reinforced soil walls. Geotext. Geomembr. 35, 25–40.
Leonardi, A., Wittel, F.K., Mendoza, M., Vetter, R., Herrmann, H.J., 2016. Particle-fluid-
increased the arrival time at the downstream area. No matter for structure interaction for debris flow impact on flexible barriers. Comput.-Aided Civil
rigid or destructible check dam, multiple check dams were favor- Infrastruct. Eng. 31 (5), 323–333.
able for hazard mitigation and it is recommended to construct the Liu, M.B., Liu, G.R., Lam, K.Y., Zong, Z., 2003. Smoothed particle hydrodynamics for
numerical simulation of underwater explosion. Comput. Mech. 30 (2), 106–118.
dams in downstream area, but the distance between check dams Liu, F.Z., Xu, Q., Dong, X.J., Yu, B., Frost, J.D., Li, H.J., 2017. Design and performance of
should not be too long. Moreover, it is recommended to construct a novel multi-function debris flow mitigation system in Wenjia Gully, Sichuan.
check dams at straight channel to avoid super-elevation. Landslides 14 (6), 2089–2104.
Luino, F., De Graff, J.V., 2012. The Stava mudflow of 19 July 1985 (Northern Italy): a
disaster that effective regulation might have prevented. Nat. Hazards Earth Syst. Sci.
Acknowledgments 12 (4), 1029–1044.
Luna, B.Q., Remaître, A., Van Asch, T.W., Malet, J.P., Van Westen, C.J., 2012. Analysis of
debris flow behavior with a one dimensional run-out model incorporating entrain-
The majority of the work described in this paper was supported by
ment. Eng. Geol. 128, 63–75.
the National Key Research and Development Program of China under Ma, G.W., An, X.M., 2008. Numerical simulation of blasting-induced rock fractures. Int. J.
Grant No. 2017YFC0804602, National Natural Science Foundation of Rock Mech. Min. Sci. 45 (6), 966–975.
Major, J.J., Pierson, T.C., 1992. Debris flow rheology: experimental analysis of fine-
China under Grant No. 41602288, the Shanghai Chenguang Scheme,
grained slurries. Water Resour. Res. 28 (3), 841–857.
the Young Elite Scientist Sponsorship Program by CAST, and the McDougall, S., Hungr, O., 2004. A model for the analysis of rapid landslide motion across
Fundamental Research Funds for the Central Universities. The authors three-dimensional terrain. Can. Geotech. J. 41 (6), 1084–1097.
would like to acknowledge all these sources of financial support and Mizuyama, T., 2008. Structural countermeasures for debris flow disasters. Int. J. Erosion
Control Eng. 1 (2), 38–43.
express the most sincere gratitude. Muramoto, Y., Uno, T., Takahashi, T., 1986. Investigation of the Collapse of the Tailings
Dam at Stava in the Northern Italy: Annuals DPRI 29A. pp. 19–52 (in Japanese).
References Murray, Y.D., Abu-Odeh, A.Y., Bligh, R.P., 2007. Evaluation of LS-DYNA Concrete
Material Model 159 (No. FHWA-HRT-05-063). Federal Highway Administration,
McLean, VA.
An, J., Tuan, C.Y., Cheeseman, B.A., Gazonas, G.A., 2011. Simulation of soil behavior Ouyang, C., Zhou, K., Xu, Q., Yin, J., Peng, D., Wang, D., Li, W., 2017. Dynamic analysis
under blast loading. Int. J. Geomech. 11 (4), 323–334. and numerical modeling of the 2015 catastrophic landslide of the construction waste
Braun, A., Cuomo, S., Petrosino, S., Wang, X., Zhang, L., 2018. Numerical SPH analysis of landfill at Guangming, Shenzhen, China. Landslides 14 (2), 705–718.
debris flow run-out and related river damming scenarios for a local case study in SW Pastor, M., Haddad, B., Sorbino, G., Cuomo, S., Drempetic, V., 2009. A depth-integrated,
China. Landslides 15 (3), 535–550. coupled SPH model for flow-like landslides and related phenomena. Int. J. Numer.
Chandler, R.J., Tosatti, G., 1995. The Stava tailings dams failure, Italy, July 1985. Proc. Anal. Methods Geomech. 33 (2), 143–172.
Instit. Civil Eng. Geotech. Eng. 113, 67–79. Pirulli, M., Barbero, M., Marchelli, M., Scavia, C., 2017. The failure of the Stava Valley
Chen, X., Cui, P., You, Y., Chen, J., Li, D., 2015. Engineering measures for debris flow tailings dams (Northern Italy): numerical analysis of the flow dynamics and rheolo-
hazard mitigation in the Wenchuan earthquake area. Eng. Geol. 194, 73–85. gical properties. Geoenviron. Disasters 4 (3), 1–15.
Chen, H.X., Zhang, L.M., 2015. EDDA 1.0: integrated simulation of debris flow erosion, Remaître, A., Van Asch, T.W., Malet, J.P., Maquaire, O., 2008. Influence of check dams on
deposition and property changes. Geosci. Model Dev. 8 (3), 829–844. debris-flow run-out intensity. Nat. Hazards Earth Syst. Sci. 8 (6), 1403–1416.
Chen, H.X., Zhang, L.M., Chang, D.S., Zhang, S., 2012. Mechanisms and runout char- Revellino, P., Hungr, O., Guadagno, F.M., Evans, S.G., 2004. Velocity and runout simu-
acteristics of the rainfall-triggered debris flow in Xiaojiagou in Sichuan Province, lation of destructive debris flows and debris avalanches in pyroclastic deposits,
China. Nat. Hazards 62 (3), 1037–1057. Campania region, Italy. Environ. Geol. 45 (3), 295–311.
Chen, H.X., Zhang, L.M., Gao, L., Yuan, Q., Lu, T., Xiang, B., Zhuang, W.L., 2017. Scheidl, C., Chiari, M., Kaitna, R., Müllegger, M., Krawtschuk, A., Zimmermann, T.,
Simulation of interactions among multiple debris flows. Landslides 14 (2), 595–615. Proske, D., 2013. Analysing debris-flow impact models, based on a small scale
Chen, H.X., Zhang, L.M., Zhang, S., 2014. Evolution of debris flow properties and physical modelling approach. Surv. Geophys. 34 (1), 121–140.
interactions in debris-flow mixtures in the Wenchuan earthquake zone. Eng. Geol. Shen, P., Zhang, L.M., Chen, H.X., Gao, L., 2017. Role of vegetation restoration in miti-
182, 136–147. gating hillslope erosion and debris flows. Eng. Geol. 216, 122–133.

61
H.-X. Chen et al. Engineering Geology 251 (2019) 48–62

Shieh, C.L., Guh, Y.R., Wang, S.Q., 2007. The application of range of variability approach flows after the 2008 Wenchuan earthquake, China. Nat. Hazards Earth Syst. Sci. 12,
to the assessment of a check dam on riverine habitat alteration. Environ. Geol. 52 (3), 201–216.
427–435. Zhang, N., Matsushima, T., 2016. Simulation of rainfall-induced debris flow considering
Song, D., Ng, C.W.W., Choi, C.E., Zhou, G.G.D., Kwan, J.S., Koo, R.C.H., 2017. Influence material entrainment. Eng. Geol. 214, 107–115.
of debris flow solid fraction on rigid barrier impact. Can. Geotech. J. 54 (10), Zhang, S., Zhang, L.M., Chen, H.X., 2014. Relationships among three repeated large-scale
1421–1434. debris flows at Pubugou Ravine in the Wenchuan earthquake zone. Can. Geotech. J.
Takahashi, T., 2007. Debris Flow: Mechanics, Prediction and Countermeasures. Taylor & 51 (9), 951–965.
Francis Group, London, UK. Zhou, G.G.D., Cui, P., Chen, H.Y., Zhu, X.H., Tang, J.B., Sun, Q.C., 2013. Experimental
Tang, C., Rengers, N.V., Van Asch, T.W., Yang, Y.H., Wang, G.F., 2011. Triggering con- study on cascading landslide dam failures by upstream flows. Landslides 10 (5),
ditions and depositional characteristics of a disastrous debris flow event in Zhouqu 633–643.
city, Gansu Province, northwestern China. Nat. Hazards Earth Syst. Sci. 11 (11), Zhou, G.G.D., Cui, P., Tang, J.B., Chen, H.Y., Zou, Q., Sun, Q.C., 2015. Experimental study
2903–2912. on the triggering mechanisms and kinematic properties of large debris flows in
Wei, R., Zeng, Q., Davies, T., Yuan, G., Wang, K., Xue, X., Yin, Q., 2018. Geohazard Wenjia Gully. Eng. Geol. 194, 52–61.
cascade and mechanism of large debris flows in Tianmo gully, SE Tibetan Plateau and Zhou, G.G.D., Song, D., Choi, C.E., Pasuto, A., Sun, Q.C., Dai, D.F., 2018. Surge impact
implications to hazard monitoring. Eng. Geol. 233, 172–182. behavior of granular flows: effects of water content. Landslides 15 (4), 695–709.
Wendeler, C., Volkwein, A., Roth, A., Denk, M., Wartmann, S., 2007. Field measurements Zollinger, F., 1985. Debris detention basins in the European Alps. In: International
and numerical modelling of flexible debris flow barriers. In: Debris-Flow Hazards Symposium on Erosion, Debris Flow and Disaster Prevention, Tsukuba, Japan, pp.
Mitig. Mech. Predict. Assess. Millpress, Rotterdam, pp. 681–687. 433–438.
Xu, Q., Zhang, S., Li, W.L., Van Asch, T.W., 2012. The 13 August 2010 catastrophic debris

62

Potrebbero piacerti anche