This paper presents the numerical results of impulsive waves generated by landslides of solid block, granular materials and heavy block sinking. An impulse product parameter P is developed and a wide range of effective parameters are studied. The volume-of-fluid (VOF) and overset mesh methods have been used to study landslide-generated tsunamis. Also, a Lagrangian tracking approach coupled with the VOF to simulate the granular movement was developed. The effect of the water reservoir depth, the landslide height, the landslide density and the geometrical parameters on the wave height (elevation) has been investigated using the open-source OpenFOAM software. The results have been presented for dimensionless distances and the normalized geometry of the landslide in the ranges 5–7, and 1–2, respectively. These numbers have been normalized to the height of the landslide (a). According to the results of simulations, the tsunami formation process is divided into three stages, which were analyzed in detail by considering the interactions between the solid and the water reservoir. The Scott Russell wave has the highest impulse product parameter among the impulse wave mechanisms which is 58.6% of the total impulse production. In addition, the duration of the wave propagation has been computed based on the wave height.

  • The impulse wave generation is modeled.

  • The importance of design parameters in the reservoir dams are highlighted.

  • The tsunami formation process is divided into three stages.

  • The novel numerical method is used to simulation.

  • Coupled Lagrangian tracking using VOF over a set mesh has been applied.

Tsunamis (impulsive waves) can be generated by sudden movements of volumes of water induced by landslides, earthquakes, volcanic eruptions, and asteroids impacts. Among these, landslides play a key role in impulsive wave generation. This phenomenon happened in the Lituya Bay in 1958 (Fritz et al. 2009), in the Vajont Valley in 1963 (Panizzo et al. 2005) and in Papua New Guinea in 1998 (Synolakis et al. 2002). Also, this phenomenon can occur in bays, reservoirs, lakes, and islands. In a dam reservoir, the behavior of the impulse wave and the density and height of landslides significantly influence the life of dams and their efficiency (Fritz et al. 2003a, 2003b, 2004).

Romano et al. (2019) numerically studied tsunamis generated by solid and impermeable landslides and proposed a novel approach to impulse wave generation. Water waves generated by landslides have been widely studied theoretically, numerically, and experimentally (Pelinovsky & Poplavsky 1996; Watts 1998; Liu et al. 2005; Panizzo et al. 2005; Tinti et al. 2005; Daneshfaraz & Kaya 2008; Synolakis et al. 2002). Heidarzadeh et al. (2020) investigated tsunamis generated by landslides that were caused by volcanic eruptions. They found that the wavelengths of landslide-generated waves are shorter than earthquake-generated waves and they exhibit greater dispersive effects. Mulligan et al. (2020) proposed a new method to simulate landslide wave generation. Their technique was an improvement upon the conventional methods that had been used. A comprehensive review of waves generated by landslides was presented by Bullard et al. (2019a, 2019b). The causes of wave generation associated with landslides were analyzed by Mulligan & Take (2017). It was found that the main criterion for the wave propagation process is the momentum transfer near the collision zone.

One of the first numerical works in this field was performed by Skvortsov & Bornhold (2007). More recently, Jing et al. (2020) developed numerical models to investigate the dispersive effects of water waves generated by an accelerated landslide. Heller (2009) used a combination of numerical, analytical, and numerical methods to study impulse waves. They analyzed landslide strength, movement of the landslide, and the wave run-up (overtopping). It was shown that, with increasing landslide strength, the wave height and run-up elevation vary. Dutykh et al. (2012) studied the flow patterns of wave propagation under the impact of a landslide.

In the recent past, scientists have proposed various water treatments (Hasegawa et al. 2019; Tom 2021), and explained conservation of water resources (Ali & Ahmad 2020; Sahoo et al. 2021), and remediating strategies for environmental (Shahid et al. 2019; Ajiboye et al. 2020). Nanotechnology has also been proposed considering efficient application (Maiga et al. 2020).

There are a few researchers who have investigated the interaction of a landslide with the free water of a dam ‫‪reservoir using numerical methods. Lindstrøm (2016) examined the influence of landslide mobility by conducting numerical simulations on porous landslides. Their results showed that the maximum near-field amplitude decreased with increasing the landslide porosity. The main objective of that work was to study the characteristics of impulsive waves generated by landslides and heavy blocks.

Shen et al. (2021) proposed a risk dynamic mechanism to overcome the flaws of safety control in reservoir dam. It was found that the proposed model can increase the abilities of reservoir dams to operate under emergency situations. Nobarinia et al. (2021) proposed a novel model to predict the peak outflow from a breached embankment dam. However, to date, studies have yet to be done to more fully investigate the mechanisms of impulse wave generation in reservoir dams. In this paper, numerical simulation has been performed to assess the effects of impulse wave generation types on the impulse product parameters P.

The rest of the paper is arranged as follows. Section 2 presents details of the mathematical modeling, the dimensionless numbers, and assumptions. Section 3 discusses the details of the geometry, the description of case studies, and details of the numerical technique. After validation of the results, Section 3–5 will present a comprehensive discussion about the results. Finally, engineering advice and design criteria are presented in Section 6.

RNAS model

The relevant equations of motion include conservation of momentum, the continuity equation, and the phase-fraction equation in their incompressible form, respectively provided by several authors (Liu et al. 2005; Ding et al. 2007; McDonough 2009; Bedram & Moosavi 2012; Ghaderi et al. 2020a, 2020b):
(1)
(2)
(3)
where denotes the gradient operator, represents the divergence operation, and is the tensor product operator. The symbols p, Fσ, and τ are the pressure, the surface tension, and the deviatoric stress tensor, respectively. The Stokes’ hypothesis and the symmetry of the stress tensor apply. Equation (3) is provided by the volume-of-fluid (VOF) model which governs the volume fraction α of one fluid in the mixture. The VOF method determines α for each cell within the computational domain. Cells having α equal to 1 contain only water and those equal to 0 contain only the gas phase (Kunkelmann & Stephan 2009).

DPM model

The discrete particle method (DPM) is adopted here to simulate the solid particle movement on the landslide by considering the effects of gravity. The trajectory of solid particles are calculated from Equation (4) (Li et al. 1999):
(4)
dp is the diameter of the solid particles and it has constant value of 5 cm. The drag force equation is adopted as follows (Li et al. 1999; Nazari et al. 2021):
(5)
where CD denotes the drag coefficient, up represents the velocity of particles, and dp is the particle diameter. In addition, because of the size of the solid particles and the interactions between them, the Brownian force (FBrownian) is considered. Details of Brownian forces can be found in Nazari et al. (2021). The movement of solid particles on the landslide and their impact with the water surface is coupled by the combination of DPM and VOF, which is a Lagrangian tracking approach.

The physical situation being considered is as follows.

Landslide model

An inclined embankment of a water body contains a mass of material (shown in Figure 1). The landslide material is triangular in shape with width b and height a. Initially, the mass is above the water surface. It moves under the influence of gravity downwards (and to the right in Figure 1). In the figure, the mass is shown after it has submerged somewhat within the water. The mass slides along the embankment until it reaches an impact location where it stops (depth H). The symbol s represents the streamwise direction of flow. The symbol h represents the initial height of the mass above the impact location. The moving mass initiates a wave (represented by the dashed line drawn a top the free surface location).

Figure 1

The physical situation with numerical parameters annotated.

Figure 1

The physical situation with numerical parameters annotated.

Close modal
In order to analyze the results, non-dimensionalization of the variables is performed using the height of the landslide a, the wide of the landslide b, the water reservoir depth H, and the landslide height h. (Ribeiro et al. 2020; see Figure 1):
(6)
where H* denotes the dimensionless depth, h* and G are the dimensionless landslide height, the geometrical ratio and streamwise, respectively.

Wave generation by landslides and turbulence effects were predicted using the Reynolds-averaged form of Equations (1)–(3). This is often called the RANS approach. The κε turbulence model is used to account for turbulence in the flow. The initial values of the turbulent kinetic energy, k and the turbulent dissipation rate, ε were 0.735 and 3.835 , respectively. For a complete description of landslide collisions with water, it is necessary to consider physical properties such as the viscosity, the density, and the surface tension coefficient. These parameters are listed in Table 1.

Table 1

Physical properties of water, air, solid particle and landslide at the STP condition

Material (kg/m3) (N/m) (Pa.s × 10−3)
Air 1.21 – – 
Water 977 0.7 0.0727 
Landslide 2,100 – – 
Solid particle 2,100 – – 
Material (kg/m3) (N/m) (Pa.s × 10−3)
Air 1.21 – – 
Water 977 0.7 0.0727 
Landslide 2,100 – – 
Solid particle 2,100 – – 
The impulse product parameter P can be found from:
(7)
where F, S and M denote the slide Froude number, the relative slide thickness and the relative slide mass, respectively. The parameter P includes all relevant slide parameters affecting the wave generation and propagation.

The open-source field operation and manipulation (OpenFOAM) CFD software package version 1912 was used to perform the numerical simulations. The OpenFOAM code is written in C ++ and uses the finite-volume discretization method to solve the conservation equations of mass and momentum, along with the equations of state. The base code of the solver is overInterDyFoam (Jasak 2009). This solver can take into account different mesh movement models. The second-order upwind scheme is used to handle the convective terms except the phase-fraction term which is discretized using the Vanleer second-order scheme. The Gauss-linear second-order approach is employed to deal with the diffusion terms. The PISO algorithm is applied to couple the pressure and the velocity components. The under-relaxation factors for the pressure, momentum, and energy equations are 0.3, 0.7, and 1, respectively. In addition, the minimum residuals for pressure, velocity, and phase fraction convergence are 10−7, 10−6, and 10−8, respectively.

Figure 1 presents a schematic view of the domain while Figure 2 shows the computational elements of the overset domain, the background domain, and the porous domain. In order to generate a high-efficiency computational mesh, a block-mesh was utilized. The computational mesh is refined gradually from the outer boundary towards the inner domain by halving the sizes of cells in a sequence of local refinement boxes. It is found that these sizes are large enough to guarantee that the results are not affected by the dimension of the domain.

Figure 2

3D view of the mesh configuration including the boundary conditions and solution domain.

Figure 2

3D view of the mesh configuration including the boundary conditions and solution domain.

Close modal

The overset mesh method is based on the use of two (or more) domains. The outer one (i.e. background domain) allows the motion of one or more inner domains (i.e. floating domains) that contain a solid body. The mutual exchange of information between the two domains is achieved by interpolation. Therefore, the two domains, which overlap each other, can be used to simulate different features of the hydrodynamics problem at hand. In Figure 2 a sketch depicting the features of the method is shown. Contrary to other techniques (e.g. immersed boundary or moving mesh methods), this method offers the great advantage that the resolution around the moving body is extremely accurate (i.e. bodyfitted approach) and remains constant throughout the simulation. Thus, the strength of the overset mesh method lies in its ability to represent complex geometries while maintaining a good quality mesh, especially for large amplitude body motions.

Due to the three-dimensional nature of this study, the computational domain is enclosed by three boundaries. No-slip and the fixed-flux pressure conditions are imposed at the landslide walls, and non-reflective-boundary conditions at the background walls and the far-field boundaries. The landslide is triangular with the dimensions a = 0.2 m and b = 0.5 m. It is initially at rest, at the distance 2 m from the reservoir bottom (Figure 3).

To validate the results of the current study, a case with landslide material density of 2,500 kg/m3 is used. Comparison of our data with the numerical and experimental visualizations presented by Liu et al. (2005) for a three-dimensional numerical simulation of the wave propagation caused by a landslide is illustrated in Figure 4. The vertical axis η refers to the water surface elevation. Figure 4 shows a comparison of the elevation of the water surface between the present work and those of Liu et al. (2005). The maximum relative deviation between the water surface elevations for the two cases in Figure 4 is ∼ 2%.

Figure 3

Schematic 3D view of the computational domain including the landslide mesh.

Figure 3

Schematic 3D view of the computational domain including the landslide mesh.

Close modal
Figure 4

Comparison of the present results with Liu et al. (2005).

Figure 4

Comparison of the present results with Liu et al. (2005).

Close modal

The numerical mesh consists of rectangular cells. These cells are refined gradually from the outer boundary toward the inner domain. A grid-independence test has been carried out to compute the proper number of numerical cells for a convergent simulation. To obtain grid independent results, simulations have been performed using three different mesh topologies and with a landslide material density of 2,500 kg/m3. The three meshes are denoted by A, B, and C. The total number of elements for the respective meshes are: A = 700,000, B = 800,000, and C = 900,000. It was observed that the B and C meshes produced almost identical results along the water surface elevation with an approximate percent error of less than 0.3%. Hence, mesh B was chosen as the preferred mesh to balance accuracy and computational time. A summary of the grid-independence test results is shown in Figure 5. The maximum skewness of the grid is 0.3611, which is suitable for obtaining accurate results.

Figure 5

Mesh independence test for water elevation at the landslide density of 2,500.

Figure 5

Mesh independence test for water elevation at the landslide density of 2,500.

Close modal

The numerical predictions were also analyzed with respect to different time-step sizes. Three different values of the time-steps are considered and the water surface elevation was computed, yielding the results presented in Figure 6. The figure shows that results are time-step independent for values of 0.00005 seconds.

Figure 6

Time-step independence test for water elevation at the landslide density of 2,500.

Figure 6

Time-step independence test for water elevation at the landslide density of 2,500.

Close modal

The process of producing a landslide-induced impulse wave involves three stages. These stages are termed as: the formation, propagation, and run-up. This paper focuses on the formation of a single wave. The landslide will move with gravity and impact with the water surface. Each case involves simulations of several positions of landslide geometry. The dimensionless ratios that were studied (h/a) are: 5, 6 and 7. The normalized geometry (a/b) of the landslide takes values of 1, 1.5 and 2.

Effect of the dimensionless distance

Figure 7 presents the time-elapsed images of the water elevation at the formation stage. Results presented in Figure 7 indicate that larger waves are near the side of the channel. This means that the generated wave starts to move along the free surface. The movement of this wave obliterates the initial shape of the free surface and leads to the propagation of smaller waves in the +x direction. Figure 7 reveals that with the increase in the landslide height (strength), the wave height increased.

Figure 7

Variation of the wave elevation at the landslide density of 2,500 kg/m3 corresponding to the three dimensionless distance cases.

Figure 7

Variation of the wave elevation at the landslide density of 2,500 kg/m3 corresponding to the three dimensionless distance cases.

Close modal

Effects of the normalized geometry of the landslide

Figure 8 shows the formation of the single wave as a function of the normalized dimension of the landslide mass. For smaller values of a/b, there is the creation of greater wave heights. The average elevation of water with normalized landslide mass geometry of 2, 1.5 and 1 and a non-dimensional landslide height of 6 are 1.15, 1.26, and 1.3, respectively. This figure shows that the height of the landslide a, has a greater influence on the wave elevation than the wide of the landslide b.

Figure 8

Variation of the wave elevation for a landslide mass density of 2,500 kg/m3 corresponding to the three normalized sizes of the mass.

Figure 8

Variation of the wave elevation for a landslide mass density of 2,500 kg/m3 corresponding to the three normalized sizes of the mass.

Close modal

In the hills overlooking the dam reservoirs, landslides with different geometry may cause impulse waves. Based on the type of these hills, the landslides with different widths can be separated. The effect of landslide geometry has also been demonstrated by Sun et al. (2020). It was found that, with the increase in the normalized geometry, the impulse wave height (wave energy) decreases.

Effect of landslide mass density

Figure 9 illustrates variation in the wave elevation with respect to time. Results show that the water elevation increases with an increase in the landslide density. The results are to be expected because more dense materials will result in a greater mass (and thus momentum) of the material. The landslide suddenly impacts and transfers its momentum to the water which is initially at rest. Subsequently, surface waves form during the sinking process.

Figure 9

Variation of the wave elevation corresponding to landslide mass density.

Figure 9

Variation of the wave elevation corresponding to landslide mass density.

Close modal

The effect of landslide mass density has also been investigated by Schuster & Highland (2003) and Djukem et al. (2020). One of the main features of momentum transferring is the disintegration of the quiescent water surface into smaller droplets at the early stages due to the gravity.

Effects of the water depth

Water depth is an important parameter for landslide-induced impulse waves. Figure 10 shows the effect of the dam water depth on the wave formation. Results show that the wave elevation increases with a decrease in the dam water depth. The reason for this is related to the effects of the bed. The landslide contact with the bed normally causes an increase in the wave height. It is important to notice that a decrease in the water depth obviously leads to more significant bed effects.

Figure 10

Variation of the wave elevation for various water-depth cases (landslide mass density = 2,500 kg/m3).

Figure 10

Variation of the wave elevation for various water-depth cases (landslide mass density = 2,500 kg/m3).

Close modal

Granular material

Granular material constituting the landslide mass was also studied. Here, the material is not cohesive and spreads as it slides along the embankment, toward the water. To begin, comparisons between the present work and prior research will be set out. In Figure 11, the results of the present study were compared with the results of Shan & Zhao (2014). The figure shows excellent agreement of water surface elevations at a particular instant in time. The normalized landslide mass geometry (a/b) is 1, and the starting location (h/a) and the density of the mass are 6 and 2,500 kg/m3, respectively. Earthen mounds around dam reservoirs are an example of granular material which may slide and created an impulse wave. The average size of the solid particles is 5 cm.

Figure 11

Comparison of water surface at a time of 0.4 s.

Figure 11

Comparison of water surface at a time of 0.4 s.

Close modal

Figure 12 shows a sequence of contours that reveal the demarcation between air and water regions. Also shown in the images are the corresponding positioning of the granular material. It can be seen that the material does not maintain a triangular shape. Rather, the granular material spreads while it slides. Disturbances in the water can be seen relatively early in the sequence of images with a small disturbance forming far from the sidewall and a larger disturbance at the impact site where the granular material enters the water.

Figure 12

VOF contours of wave propagation corresponding to a granular landslide at same conditions as the solid mass.

Figure 12

VOF contours of wave propagation corresponding to a granular landslide at same conditions as the solid mass.

Close modal

Next, attention is turned toward waves generated by a heavy rectangular cross-sectional object sinking vertically into water, again using the VOF and the overset mesh methods. This configuration is known as a solitary Scott Russell wave formation in a long rectangular tank. A combination of VOF and overset mesh was used to clarify details of the solitary wave formation and its propagation. First, we simulated a long rectangular tank with a heavy/dense rectangular block adjacent to the left side, as shown in Figure 13. The generated wave profile of the present work is compared with experimental observations of Ataie-Ashtiani & Shobeyri (2008) (upper case of Figure 13) and Monaghan & Kos (2000) (lower case of Figure 13). Based on the excellent agreement between the present calculations and two prior references, confidence is provided for the current calculations. With this verification completed, various new scenarios will be simulated and discussed. The density of sinking box is 2.5 times the water density and the initial height above the water.

Figure 13

Formation of Scott Russell wave propagation, a comparison between present results and prior research.

Figure 13

Formation of Scott Russell wave propagation, a comparison between present results and prior research.

Close modal

Effects of the heavy block density

The first issue to consider is the influence of block density on the wave formation. A series of calculations with density ranging from 1,800 to 3,300 kg/m3 was performed and the results are shown in Figure 14. It is seen that a lower density block reduces the magnitude of the wave profile. It is well known that single wave profiles in long tanks significantly depend on block density, whereas generally the wave elevation increases with density. The characteristics of the wave is similar to the landslide-induced wave discussed earlier. In Figure 14, wave profiles with different block densities are depicted for the Scott Russell wave case study. The density with 1,800 kg/m3 exhibits the smallest wave profile peak of approximately 0.295 m. The wave profile peaks increase with density (see Figure 15).

Figure 14

Variation of the wave profile corresponding to four densities of the block at 0.7 s.

Figure 14

Variation of the wave profile corresponding to four densities of the block at 0.7 s.

Close modal
Figure 15

Variation of the VOF contours of wave propagation corresponding to four densities of the block at 0.7 s.

Figure 15

Variation of the VOF contours of wave propagation corresponding to four densities of the block at 0.7 s.

Close modal

Effect of the water depth

Five water depths were investigated with a constant block density of 2,500 kg/m3. The resulting wave profiles are illustrated in Figure 16. During the sinking, the peak of the Scott Russell wave has a maximum velocity of 10.5 m/s, which decreases gradually. Figures 16 and 17 show the effect of the water depth on the wave formation at time 0.7 s. Results show that the water elevation increases with decreasing water depth.

Figure 16

Variation of wave profile for a landslide density of 2,500 kg/m3 corresponding to the five water-depth cases, results obtained at 0.7 s.

Figure 16

Variation of wave profile for a landslide density of 2,500 kg/m3 corresponding to the five water-depth cases, results obtained at 0.7 s.

Close modal
Figure 17

V wave propagation corresponding to the five water depth cases, results correspond to 0.7 s.

Figure 17

V wave propagation corresponding to the five water depth cases, results correspond to 0.7 s.

Close modal

Effects of the block height

A free-falling dense block that impacts a water surface creates a single wave which differs from a landslide wave. The elevation of the wave is greater than the landslide wave due to the greater energy transfer to the water (compare Figures 7 and 18). The impact velocity of the heavy block, when friction is neglected, can be calculated using:
(8)
Figure 18

Wave profiles for a landslide density of 2,500 kg/m3 corresponding to the four block heights, results obtained at 0.7 s.

Figure 18

Wave profiles for a landslide density of 2,500 kg/m3 corresponding to the four block heights, results obtained at 0.7 s.

Close modal
For a block starting from rest, the impact velocity is:
(9)
where Δh and g are the block height and gravitational acceleration, respectively. It was found that blocks with a small height create a minimal solitary wave due to their lower impact momentum. Results of Figures 18 and 19 indicate that with an increase in the block height, the wave profile increases because of the increased mass (and thus momentum) of the block upon impact. An increasing box height causes an increase of the momentum. The wave height increases with increasing momentum. Figure 19 discloses water ascension rates corresponding to the four block heights. Results are displayed at a time of 0.7 s. The crest of the wave increases from 0.29 m to 0.34 m as the box height increases from 0.4 m to 0.55 m. Also, the crest profile is stretched along the vertical direction. Behaviors of the crest profile significantly affect the vortex under the wave.
Figure 19

Wave propagation corresponding to the four block height cases at 0.7 s.

Figure 19

Wave propagation corresponding to the four block height cases at 0.7 s.

Close modal

The impulse product parameter P applies to landslides and to granular and block impulse waves. Numerical results based on P indicate that the largest waves occur for a block sinking into the water (Figure 20). These large waves are a negative phenomenon, particularly for dam reservoirs.

Figure 20

Comparison of impulse product parameter for all of three case studies.

Figure 20

Comparison of impulse product parameter for all of three case studies.

Close modal

The main focus of the present paper is on the generation of landslide-induced impulse waves and Scott Russell waves. Three different positions have been selected based on the simulations conducted for a VOF and overset mesh using the OpenFOAM C ++ library. Three-dimensional simulations of the impulse wave formation for landslide densities ranging from 2,100 to 2,900 kg/m3 and Scott Russell waves for densities ranging from 1,800 to 3,300 kg/m3 were carried out. It is found that the formation of water waves corresponding to a landslide density of 2,500 kg/m3 is consistent with recent experimental results. The present study opens exciting possibilities for future research relating to the design of dam reservoir. Some conclusions about the different features of the wave generation are summarized as follows.

  • 1. The coupling the overset mesh technique overcomes a drawback of the overset mesh method as far as the modelling of a solid body moving in contact with an impermeable surface. The proposed numerical method can be used for Scott Russell wave generation.

  • 2. The landslide-induced impulse wave process is characterized by three stages. In this study, we focused on the first stage; however, the three stages are:

    • - Formation of the single wave,

    • - Propagation of the wave,

    • - Run-up (overtopping).

  • 3. With the increase in the normalized geometry, the impulse wave height decreases.

  • 4. With the increase in the box height, the impulse wave height increases.

  • 5. With the increase in the water depth, the impulse wave height decreases.

  • 6. With the increase in the heavy block density, the impulse wave height increases.

  • 7. The impulse product parameter of the Scott Russell wave is calculated to be 4.98. This value for the landslide and granular are 3 and 1.89, respectively.

As the landslide density and the dimensionless distance of the landslide increase, the magnitude of the wave increases. Also, with the decrease in water depth, the wave elevation increases.

All relevant data are included in the paper or its Supplementary Information.

Ajiboye
T. O.
,
Kuvarega
A. T.
&
Onwudiwe
D. C.
2020
Recent strategies for environmental remediation of organochlorine pesticides
.
Applied Sciences
10
(
18
),
6286
.
Ali
Z.
&
Ahmad
R.
2020
Nanotechnology for water treatment
. In:
Environmental Nanotechnology
, Vol.
3
(N. Dasgupta, S. Ranjan, E. Lichtfouse, eds).
Springer
,
Cham
, pp.
143
163
.
Ataie-Ashtiani
B.
&
Shobeyri
G.
2008
Numerical simulation of landslide impulsive waves by incompressible smoothed particle hydrodynamics
.
International Journal for Numerical Methods in Fluids
56
,
209
232
.
Bedram
A.
&
Moosavi
A.
2012
Breakup of droplets in micro and nanofluidic T-junctions
. In:
Applied Mechanics and Materials
, Vol.
110
(Wu Fan, ed.).
Trans Tech Publications Ltd
, pp.
3673
3678
.
Bullard
G. K.
,
Mulligan
R. P.
,
Carreira
A.
&
Take
W. A.
2019a
Experimental analysis of tsunamis generated by the impact of landslides with high mobility
.
Coastal Engineering
152
,
103538
.
Bullard
G. K.
,
Mulligan
R. P.
&
Take
W. A.
2019b
An enhanced framework to quantify the shape of impulse waves using asymmetry
.
Journal of Geophysical Research: Oceans
124
(
1
),
652
666
.
Daneshfaraz
R.
&
Kaya
B.
2008
Solution of the propagation of the waves in open channels by the transfer matrix method
.
Ocean Engineering
35
(
11–12
),
1075
1079
.
Ding
H.
,
Spelt
P. D.
&
Shu
C.
2007
Diffuse interface model for incompressible two-phase flows with large density ratios
.
Journal of Computational Physics
226
(
2
),
2078
2095
.
Djukem
W. D. L.
,
Braun
A.
,
Wouatong
A. S. L.
,
Guedjeo
C.
,
Dohmen
K.
,
Wotchoko
P.
,
Fernandez-Steeger
T. M.
&
Havenith
H. B.
2020
Effect of soil geomechanical properties and geo-environmental factors on landslide predisposition at Mount Oku, Cameroon
.
International Journal of Environmental Research and Public Health
17
(
18
),
6795
.
Dutykh
D.
,
Mitsotakis
D.
,
Beisel
S. A.
&
Shokina
N. Y.
2012
On waves generated by an underwater landslide
. In:
IV Conference ‘Applied Problems of the Fluid Mechanics, Heat and Mass Transfer’
. pp.
75
79
.
Fritz
H. M.
,
Hager
W. H.
&
Minor
H. E.
2003a
Landslide generated impulse waves
.
Experiments in Fluids
35
(
6
),
505
519
.
Fritz
H. M.
,
Hager
W. H.
&
Minor
H. E.
2003b
Landslide generated impulse waves. 2
.
Hydrodynamic Impact Craters. Experiments in Fluids
35
(
6
),
520
532
.
Fritz
H. M.
,
Hager
W. H.
&
Minor
H. E.
2004
Near field characteristics of landslide generated impulse waves
.
Journal of Waterway, Port, Coastal, and Ocean Engineering
130
(
6
),
287
302
.
Fritz
H. M.
,
Mohammed
F.
&
Yoo
J.
2009
Lituya Bay landslide impact generated megatsunami 50th anniversary
.
Pure and Applied Geophysics
166
(
1–2
),
153
175
.
Ghaderi
A.
,
Daneshfaraz
R.
,
Abbasi
S.
&
Abraham
J.
2020a
Numerical analysis of the hydraulic characteristics of modified labyrinth weirs
.
International Journal of Energy and Water Resources
4
(
4
),
425
436
.
Ghaderi
A.
,
Daneshfaraz
R.
,
Dasineh
M.
&
Di Francesco
S.
2020b
Energy dissipation and hydraulics of flow over trapezoidal–triangular labyrinth weirs
.
Water
12
(
7
),
1992
.
Hasegawa
S.
,
Tanaka
Y.
,
Wake
N.
,
Takagi
R.
&
Matsuyama
H.
2019
Improving chemical cleaning of fouled membranes in a drinking water treatment plant
.
Water Supply
19
(
8
),
2330
2337
.
Heidarzadeh
M.
,
Ishibe
T.
,
Sandanbata
O.
,
Muhari
A.
&
Wijanarto
A. B.
2020
Numerical modeling of the subaerial landslide source of the 22 December 2018 Anak Krakatoa volcanic tsunami, Indonesia
.
Ocean Engineering
195
,
106733
.
Heller
V.
2009
Landslide generated impulse waves experimental results
. Coastal Engineering 2008 - 31st International Conference, Vol.
5
. pp.
1313
1325
.
Jasak
H.
2009
OpenFOAM: open source CFD in research and industry
.
International Journal of Naval Architecture and Ocean Engineering
1
(
2
),
89
94
.
Jing
H.
,
Gao
Y.
,
Liu
C.
&
Hou
J.
2020
Far-field characteristics of linear water waves generated by a submerged landslide over a flat seabed
.
Journal of Marine Science and Engineering
8
(
3
),
196
.
Kunkelmann
C.
&
Stephan
P.
2009
CFD simulation of boiling flows using the volume-of-fluid method within OpenFOAM
.
Numerical Heat Transfer, Part A: Applications
56
(
8
),
631
646
.
Lindstrøm
E. K.
2016
Waves generated by subaerial slides with various porosities
.
Coastal Engineering
116
,
170
179
.
Liu
P. F.
,
Wu
T. R.
,
Raichlen
F.
,
Synolakis
C. E.
&
Borrero
J. C.
2005
Runup and rundown generated by three-dimensional sliding masses
.
Journal of Fluid Mechanics
536
(
1
),
107
144
.
Maiga
D. T.
,
Mamba
B. B.
&
Msagati
T. A.
2020
Distribution profile of titanium dioxide nanoparticles in South African aquatic systems
.
Water Supply
20
(
2
),
516
528
.
McDonough
J. M.
2009
Lectures in Elementary Fluid Dynamics: Physics, Mathematics and Applications
.
Monaghan
J. J.
&
Kos
A.
2000
Scott Russell's wave generator
.
Physics of Fluids
12
,
622
630
.
Mulligan
R. P.
&
Take
W. A.
2017
On the transfer of momentum from a granular landslide to a water wave
.
Coastal Engineering
125
,
16
22
.
Mulligan
R. P.
,
Franci
A.
,
Celigueta
M. A.
&
Take
W. A.
2020
Simulations of landslide wave generation and propagation using the particle finite element method
.
Journal of Geophysical Research: Oceans
125
(
6
),
e2019JC015873
.
Nazari
A.
,
Jafari
M.
,
Rezaei
N.
,
Taghizadeh-Hesary
F.
&
Taghizadeh-Hesary
F.
2021
Jet fans in the underground car parking areas and virus transmission
.
Physics of Fluids
33
(
1
),
013603
.
Nobarinia
M.
,
Kalateh
F.
,
Nourani
V.
&
Amini
A. B.
2021
Dam failure peak outflow prediction through GEP-SVM meta models and uncertainty analysis
.
Water Supply
. in press.
Panizzo
A.
,
De Girolamo
P.
,
Di Risio
M.
,
Maistri
A.
&
Petaccia
A.
2005
Great Landslide Events in Italian Artificial Reservoirs
.
Pelinovsky
E.
&
Poplavsky
A.
1996
Simplified model of tsunami generation by submarine landslides
.
Physics and Chemistry of the Earth
21
(
1–2
),
13
17
.
Ribeiro
H. D. B.
,
Simões
A. L. A.
,
da Luz
L. D.
,
Mangieri
L. S. G.
&
Schulz
H. E.
2020
Stability of Solids in Stepped Flume Nappe Flows: Subsidies for Human Stability in Flows
.
Romano
A.
,
Lara
J.
,
Barajas
G.
,
Di Paolo
B.
,
Bellotti
G.
,
Di Risio
M.
,
Losada
I.
&
De Girolamo
P.
2019
Numerical modelling of Landslide-Generated Tsunamis with OpenFOAM®: a new approach
.
Coastal Structures
2019
,
486
495
.
Sahoo
T.
,
Sahu
J. R.
,
Panda
J.
,
Hembram
M.
,
Sahoo
S. K.
&
Sahu
R.
2021
Nanotechnology: an efficient technique of contaminated water treatment
. In:
Contaminants in Drinking and Wastewater Sources
(M. Kumar, D. Snow, R. Honda, S. Mukherjee, eds).
Springer
,
Singapore
, pp.
251
270
.
Schuster
R. L.
&
Highland
L. M.
2003
Impact of landslides and innovative landslide-mitigation measures on the natural environment
. In:
International Conference on Slope Engineering
,
Hong Kong, China
, Vol.
8
, No.
10
.
Shahid
M. J.
,
Tahseen
R.
,
Siddique
M.
,
Ali
S.
,
Iqbal
S.
&
Afzal
M.
2019
Remediation of polluted river water by floating treatment wetlands
.
Water Supply
19
(
3
),
967
977
.
Shen
G.
,
Lu
Y.
,
Zhang
S.
,
Xiang
Y.
,
Sheng
J.
,
Fu
J.
&
Liu
M.
2021
Risk dynamics modelling of reservoir dam break for safety control in the emergency response process
.
Water Supply
21
(
3
),
1356
1371
.
Skvortsov
A.
&
Bornhold
B.
2007
Numerical simulation of the landslide-generated tsunami in Kitimat Arm, British Columbia, Canada, 27 April 1975
.
Journal of Geophysical Research: Earth Surface
112
,
F02028
.
Sun
J.
,
Wang
Y.
,
Huang
C.
,
Wang
W.
,
Wang
H.
&
Zhao
E.
2020
Numerical investigation on generation and propagation characteristics of offshore tsunami wave under landslide
.
Applied Sciences
10
(
16
),
5579
.
Synolakis
C. E.
,
Bardet
J. P.
,
Borrero
J. C.
,
Davies
H. L.
,
Okal
E. A.
,
Silver
E. A.
,
Sweet
S.
&
Tappin
D. R.
2002
The slump origin of the 1998 Papua New Guinea tsunami
.
Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences
458
(
2020
),
763
789
.
Tinti
S.
,
Manucci
A.
,
Pagnoni
G.
,
Armigliato
A.
&
Zaniboni
F.
2005
The 30 December 2002 Landslide-Induced Tsunamis in Stromboli: Sequence of the Events Reconstructed From the Eyewitness Accounts
.
Tom
A. P.
2021
Nanotechnology for sustainable water treatment–a review
.
Materials Today: Proceedings
. in press.
Watts
P.
1998
Wavemaker curves for tsunamis generated by underwater landslides
.
Journal of Waterway, Port, Coastal, and Ocean Engineering
124
(
3
),
127
137
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).