The present study numerically investigated the liquid-gas-particles mixture flow during dam break using the computational fluid dynamics (CFD) and discrete element method (DEM). The adopted numerical methods were validated through comparison with previous researches of computation and experiments. The focus is on the effect of the particle friction coefficient (PFC) and the ratio of a particle size to a baseline size (PS) on behavior of the particles and the free surface. The PFC ranges from 0.1 to 0.9 and the PS varies from 50% to 100%. The liquid front position and the formation of the free-surface are almost independent of PFC and PS. In contrast, the particle front position strongly depends on PFC and PS. The particle front velocity showed two pattern according to time. First pattern is the increasing and then saturated behavior. Second pattern has increasing and decreasing behavior. These two patterns are mainly governed by the PFC. Generally, a smaller PFC and a larger PFC lead to first and second patterns, respectively. We could classify all cases of PFC and PS by the arrival time of particles into earlier and later arrivals than liquid front.

The presence of multiple phases such as liquid, gas, and particles can influence the behavior of a flow due to their interaction. The interaction is naturally complex and nonlinear, which makes the investigation of liquid-gas-particle flows extremely challenging. Nevertheless, these types of flows have attracted numerous researchers because they are widely observed and play an important role in various industries and nature.1–17 

It is difficult to study such flows through field observations and simple physical experiments. Therefore, most studies on these flows make significant simplifications, but they include the essential material properties and boundary conditions.18 In this view, the liquid-gas dam break flows with particles can be regarded as a simplified problem that involves essential elements of complex multiphase flows.

The numerical studies related to the liquid-gas-particle flows in dam breaks have mainly developed and validated numerical methods of such flows. Few researchers have investigated the behavior and characteristics of liquid-gas-particle flows. Guo and Morita17 compared the behavior of liquid-gas-particle flows according to height of the particle bed. They showed that the front position of the liquid became faster as the particle bed became higher. Xia et al.19 and Ran et al.20 emphasized that the sediment bed scouring and grain movement in a dam break are interesting subjects for researchers. Therefore, they investigated the propagation and physics of scouring near the interface between particles and the liquid.

Lin and Chen21 investigated the effect of the number of particles on the behavior of liquid-gas-particle flows and the impulse on a structure. The flow speed increased with decreasing concentration, while the impulse decreased. In addition, they examined the effects of parameters such as the stiffness, damping ratio, and friction coefficient of particles. The results showed that the friction coefficient has an obvious influence on the flow, while the stiffness and damping ratio did not. However, the parameters considered in that study were limited, and only the front position of the mixture flow was compared to investigate their effects.

Many numerical models for the computational fluid dynamics (CFD) have been developed for fluid flow with a free surface based on Eulerian or Lagrangian frameworks. For particle motion, the discrete element method (DEM) calculates the collision forces between particles and has been used to capture their motion.22 This method has been utilized by some researchers to model particle motion.23,24 Sun and Sakai25 and Guo et al.26 successively computed the behavior of a liquid-gas-particle mixture flow by adopting CFD-DEM which was validated by comparing with results of the experiments.

Also, the fluid flow has been simulated as particle motion in a Lagrangian framework using smoothed-particle hydrodynamics (SPH)27 and the moving particle semi-implicit (MPS) method.28 For instance, the sloshing was studied by using MPS-DEM which realized the liquid and the particles, respectively (Guo and Morita17). The SPH-DEM was used to numerically simulate two phase flows with liquid-particle for a dam break and a rotating tank.29 

In this study, we used CFD-DEM to simulate the liquid-gas-particle flow because recent studies showed reasonable results using this method.25,26 The numerical method used in this study consisted of OpenFOAM for CFD30 and CFDEMproject.31,32 The present numerical methods were successively used in our previous study for the multiphase flow of a liquid-gas-particle mixture in dam break.33 

Previous studies related to liquid-gas-particle flows have not been thoroughly discussed compared to liquid-gas flows in dam breaks.25 Since the general particles in a granular flow have different characteristics and shapes, they can affect the flow. Their coefficients of friction and elastic restitution may affect the flow behavior due to interactions with other particles or the walls.34 In addition, the effect of grain size on the hydraulic permeability is also an important characteristic of particles.18 In this sense that particle’s properties have strong influence on the multiphase flow, we had investigated dependence of multiphase flow on particle density.33 

Park et al.33 studied the effect of particle density on three phases of liquid-gas-particle flow which occurs in a dam break. The adopted numerical method of CFD-DEM successively realized the interaction of two-phase fluids and particle. They systematically analyzed the dependence of the front head speeds of the free-surface and particles on the density of the particle. They reported that the arrival time of the front head of the free surface to the wall is almost the same for all particle densities. Otherwise, the arrival time of the front head of particles becomes shorter with increasing the particle density. The decrease rate of the arrival time of the front head of particles is about the linear relation to the particle density. Finally, they provided the motion map.

Authors’ previous study33 considered only one particle’s property, finding the effect of the particle density on the complex multiphase flow of liquid-gas-particle mixture in a dam-break. The present study extents our previous study33 by considering more particle’s properties. Therefore, the purpose of the present study investigates the effects of the particles’ friction coefficient and size on their behaviors and the free surface when the mixture moves in a dam break problem. We use numerical methods to simulate the liquid-gas-particle flow. This may obtain more useful information than experimental approaches because they are still limited in observing complicated phenomena.25 The governing equations used in this study are presented in Section II. The behavior of particles and the free surface with different particle friction coefficients and diameters are compared in Section III to investigate their effects.

The present study uses the same numerical methods as our previous paper of Park et al.33 which provides a very detailed explanation for numerical methods and used OpenFOAM. Therefore, we very concisely describe the present numerical methods.

The motion of particles is realized by DEM. The translational motion is governed by the following equation:

midvidt=jFn,ij+mig+Ffluid(Fd+Fp)
(1)

where the subscripts of i and j identify the particles. Thus, mi and vi are a mass and velocity of particle i, respectively. Fn,ij, g, Ffluid (Ffluid = Fd + Fp) are the normal contact force, gravity, and the fluid force, fluid drag force, and fluid pressure gradient force, respectively.

The governing equation of the rotational motion of particles is as follows;

Iidωidt=Ri×Ft,ij
(2)

where Ii, ωi, Ri, and Ft,ij are the moment of inertia, angular velocity, distance from the center of mass to the contact point, and contact force in the tangential direction, respectively. The definitions of each forces in equations (1) and (2) can be refereed in Park et al.33 Note, the PFC (particle friction coefficient) is a key parameter in this study. The present PFC is the static friction coefficient which is defined as the ratio of the tangential force (Ft,ij) to the normal force (Fn,ij).33 In this study, the static friction coefficient is symbolized as μS which corresponds to Cfs in Park et al.33 

The continuity and the momentum equations govern the present unsteady viscous three-dimensional fluid flow as follows;

εuixi=0
(3)
tρεui+xjερuiuj=εpxi+xjεμuixj+ujxi+ερgiSm
(4)

where xi, ui, and ε are the is a Cartesian coordinate corresponding velocity component, and the porosity, respectively. In equation (4), p, ρ, and μ are the pressure, the density, and the viscosity, respectively. The two-way coupling in fluids and particles is performed by the source term Sm in equation (4).

The VOF method is utilized to track the free surfaces as the interface between the liquid and gas. The free surface in each cell is defined when a volume fraction (Q) corresponding to each fluid is between 0 and 1. The transport of Q is governed by following equation;

εQt+εuiQxi=0
(5)

The porosity (ε) in equation (5) is considered in the conventional VOF equation, which represents the fluid displaced by particles.35 The porosity is a fraction of void volume over the total volume.

The spatial discretization for the convection and diffusion terms are carried out by using a second-order upwind scheme and second-order central differencing scheme, respectively. The temporal discretization of the time derivative terms are performed by a second-order accurate backward implicit scheme. The PISO (Pressure Implicit with Splitting of Operators) algorithm36 is used for the velocity–pressure coupling and overall solution procedure. In addition, the schematics of a flow chart for the algorithm of the CFD-DEM can be refereed in Park et al.33 

Figure 1 shows that the three-dimensional computational domain is composed of a length, a breadth and a height (L, B, H) with dimension of (0.2 m, 0.1 m, 0.3 m). The initial liquid column is formed with a length, a breadth and a height (Ll, Bl, Hl) with dimension of (0.05 m, 0.1 m, 0.1 m). The number of particles is 3,883 which is stacked on the left side bottom. The air occupies the remained space. The properties of the particles and fluids are recorded in Table I and Table II, respectively.

FIG. 1.

Computational domain.

FIG. 1.

Computational domain.

Close modal
TABLE I.

Properties of particles.

TypeValue
Baseline size, do (m) 0.0027 
Diameter, dp (m) 0.00135, 0.00162, 0.00189, 
 0.00216, 0.00243, 0.0027 
Density, ρp (kg/m32,500 
The number of particles 3,883 
Spring constant (N/m) 1,000 
Restitution coefficient 0.9 
Friction coefficient, μs 0.1, 0.2, 0.3, 0.4, 0.5, 
 0.6, 0.7, 0.8, 0.9 
TypeValue
Baseline size, do (m) 0.0027 
Diameter, dp (m) 0.00135, 0.00162, 0.00189, 
 0.00216, 0.00243, 0.0027 
Density, ρp (kg/m32,500 
The number of particles 3,883 
Spring constant (N/m) 1,000 
Restitution coefficient 0.9 
Friction coefficient, μs 0.1, 0.2, 0.3, 0.4, 0.5, 
 0.6, 0.7, 0.8, 0.9 
TABLE II.

Properties of fluids.

TypeValue
Liquid density, ρl (kg/m31,000 
Liquid viscosity (Pa·s) 10-3 
Gas density, ρg (kg/m3
Gas viscosity (Pa·s) 10-5 
TypeValue
Liquid density, ρl (kg/m31,000 
Liquid viscosity (Pa·s) 10-3 
Gas density, ρg (kg/m3
Gas viscosity (Pa·s) 10-5 

For the case of a dam break flow, the Reynolds number is generally defined as Re=HgHρf/μf where ρf, μf and g are the density and viscosity of fluid, and the gravity, respectively. This definition of Re leads to about 1.0×105 as the value of the present Reynolds number. Like our previous study,33 we concentrate on the fluid flow and behavior of particles on a dry bed. From the collapse to the collision to the opposite wall, the dam break flow on a dry bed does not exhibit strong turbulence.37 Thus, this period on a dry bed is mainly exposed to the laminar flow rather than the turbulence. Therefore, the turbulence phenomena and turbulence model are not considered in this study.

In order to evaluate the drag force action on each particles in DEM, the particle Reynolds number (Rep) is needed.33 The definition of Rep is Rep=εdpvfvpρf/μf, where dp, vf and vp are the particle diameter, fluid velocity and the particle velocity, respectively. The values of the present Rep according to the particle diameter are in the range of about 1-4×102.

The structured grid is used and the grid size for the fluid flow is bigger than the particle size to avoid numerical errors in CFD-DEM. The vertical and horizontal walls of the computational domain were considered to have no-slip conditions. The dimensionless parameters to identify the effect of the particle friction coefficient and diameter are achieved by the characteristics length and time of Ll and 2g/Ll12, respectively.

The validation of the present numerical method has been perform by comparing the present front head positions of the free surface and the particles with the previous computational and experimental results of Sun and Sakai.25Figure 2 shows that the present results give a great agreement with both numerical and experimental results, which supports the reliability of the present numerical methods. In addition, the qualitative comparisons are shown in Fig. 3 where the time evolution of the formation of the free-surface and particles is presented with that of Sun and Sakai.25 The thick blue line indicates the free surface. The particles are colored individually with their velocity.

FIG. 2.

Comparison of the present dimensionless front and height positions of liquid with computation (Com) and experiment (Exp) results of Sun and Sakai.25 

FIG. 2.

Comparison of the present dimensionless front and height positions of liquid with computation (Com) and experiment (Exp) results of Sun and Sakai.25 

Close modal
FIG. 3.

Comparison between Sun and Sakai’s results25 (left column) and the present results (right column) for the particle distribution and free-surface formation at (a) t*=2, (b) t*=4 and (c) t*=6.

FIG. 3.

Comparison between Sun and Sakai’s results25 (left column) and the present results (right column) for the particle distribution and free-surface formation at (a) t*=2, (b) t*=4 and (c) t*=6.

Close modal

Sun and Sakai25 considered the effects of a rising gate by adding an obstacle in front of the liquid column. However, the present study ignores the gate effect because the focus is on the mixture flow. The overall process of the free surface and particle motion was similar to that of Sun and Sakai25 except for some differences in the free surface in the initial stages. After the collapse of the liquid column, the free surface advances against the wall, the particle bed expands over the bottom of the domain, and the particle front moves more slowly than the liquid front at t*< 4 in Figs. 3(a) and (b) Even after hitting the opposite wall, the splash of the particles and the upward motion of the free surface are reasonably represented by the present results at t*= 6 in Fig. 3(c).

We identified the effects of the particle friction coefficients (PFC) and particle sizes (PS) on the three-phase liquid-gas-particle flow. Figures 4 and 5 show typical three-dimensional perspective views and side views of the particles and the free surface between the liquid and the gas for different PFCs and PSs at t*=1.0, 2.0, 3.0, and 3.5. In these cases, PFC=0.1 and 0.9, and PS=0.5 and 1.0.

FIG. 4.

Perspective view of dam break flow with particles for (a) dp=0.5do & μs=0.1, (b) dp=0.5do & μs=0.9, (c) dp=1.0do & μs=0.1 and (d) dp=1.0do & μs=0.9 at t*=1 (1st column), t*=2 (2nd column), t*=3 (3rd column) and t*=3.5 (4th column).

FIG. 4.

Perspective view of dam break flow with particles for (a) dp=0.5do & μs=0.1, (b) dp=0.5do & μs=0.9, (c) dp=1.0do & μs=0.1 and (d) dp=1.0do & μs=0.9 at t*=1 (1st column), t*=2 (2nd column), t*=3 (3rd column) and t*=3.5 (4th column).

Close modal
FIG. 5.

Side view of dam break flow with particles for (a) dp=0.5do & μs=0.1, (b) dp=0.5do & μs=0.9, (c) dp=1.0do & μs=0.1 and (d) dp=1.0do & μs=0.9 at t*=1 (1st column), t*=2 (2nd column), t*=3 (3rd column) and t*=3.5 (4th column).

FIG. 5.

Side view of dam break flow with particles for (a) dp=0.5do & μs=0.1, (b) dp=0.5do & μs=0.9, (c) dp=1.0do & μs=0.1 and (d) dp=1.0do & μs=0.9 at t*=1 (1st column), t*=2 (2nd column), t*=3 (3rd column) and t*=3.5 (4th column).

Close modal

When the liquid column starts to collapse due to gravity, the front of the liquid and particles advances against the wall. At a fixed time, the liquid front position and the formation of the free-surface are almost independent of the variation of PFC and PS, which is clarified by the side views in Fig. 5. This independent behavior of the liquid front position and the formation of the free-surface can be also founded in Park et al.33 where the variation of the particle density is considered to investigate a multiphase liquid-gas-particle flow from a dam break using CFD-DEM.

In contrast, the particle front positions differ according to PFC and PS at a fixed time, which can be clarified by comparing the liquid front independently of PFC and PS. Briefly, a smaller PFC contributes to the fast advance of particles when PS is the same, as shown in Fig. 4. A smaller particle advances quickly when the PFC is the same, as shown in the figure. Eventually, as both PFC and PS become smaller, the particles become faster and finally reach the wall earlier, as shown in Fig. 4 and Fig. 5. Particularly, the largest PS and PFC lead to slower travel of the particle front than the liquid front up to the wall, as shown in Fig. 4(d) and Fig. 5(d).

To clarify the time evolution of the front position of the liquid and particles, the spanwise-averaged dimensionless front position of the liquid and particles (<xl*> and <xp*>) along the time are presented in Fig. 6. Additionally, only two phases of liquid-gas are considered and then a dam break flow without particles is numerically simulated to identify the effect of the particles on the dam break flow by comparing with those of the liquid-gas-particle flow.

FIG. 6.

Time histories of dimensionless front position of (a) liquid and (b) particles for different particle diameters (dP) and particle friction coefficients (μs). Here, w/o and w/particle identify the liquid-gas flow without particles and the liquid-gas-particle flow, respectively.

FIG. 6.

Time histories of dimensionless front position of (a) liquid and (b) particles for different particle diameters (dP) and particle friction coefficients (μs). Here, w/o and w/particle identify the liquid-gas flow without particles and the liquid-gas-particle flow, respectively.

Close modal

In general, the liquid front in the liquid-gas-particle flows advances slower than that in the liquid-gas flow against the wall. Irrespective of the existence of particles, the liquid front position is almost the same at the beginning of the dam break flow, but the particles give an effect on the front head position of the liquid after t*=0.5. As time goes by, the difference in the liquid front position increases until the liquid front hits the opposite wall, as observed in Fig. 6(a).

Irrespective of PFC and PS, the profile of the liquid front position in the liquid-gas-particle flows are almost the same, as shown in Fig. 6(a), which confirms the front formations of the liquid in Figs. 4 and 5. As a result, it can be concluded that the PFC and PS give a negligible effect on the liquid front.

The effect of PFC and PS on the time evolution of the particle front position is identified in Fig. 6(b). In order to direct compare the liquid front head, the positions of liquid front head without and with particles are put together with the particle front positions in Fig. 6(b). The particles move faster when simultaneously decreasing the size and friction coefficient. Thus, the smallest values of PS=0.5 and PFC= 0.1 lead to the earliest arrival of the particle front at the opposite wall. Particularly, this particle front runs ahead of the liquid front at all times until the impingement. The largest values of PS=1.0 and PFC=0.9 lead to the slowest advancing speed of the particle front over the time. Consequently, the arrival time of the particle front at the opposite wall is strongly dependent on PS and PFC. This will be discussed later for Fig. 11 where we cover all ranges of PS and PFC and classify the cases of PS and PFC into earlier and later arrivals of the particle front than the liquid front head.

Recently, authors’ previous study33 considered the effect of the particle density on the multiphase liquid-gas-particle flow from a dam break using CFD-DEM. We showed almost the invariance of the liquid front position and the formation of the free-surface according to the particle density. However, the particle front positions strongly depended on the particle density. The PFC and PS also gives the same effect on the liquid front position, the formation of the free-surface, and the particle front position as the particles density.

Figure 7 presents the dimensionless velocity and acceleration of the particle front for different PFC and PS. In addition, the liquid fronts in the liquid-gas and the liquid-gas-particle flows are considered in Fig. 7. When particles are present, the liquid front velocity reaches the maximum value at t*=2.58. The time obtained without particles is about 2.5, as shown in Fig. 7(a).

FIG. 7.

Time histories of dimensionless (a) velocity and (b) acceleration for different particle diameters (dP) and particle friction coefficients (μs).

FIG. 7.

Time histories of dimensionless (a) velocity and (b) acceleration for different particle diameters (dP) and particle friction coefficients (μs).

Close modal

The velocity of the liquid front without particles is always faster than that of particles until the liquid front velocity reached the maximum velocity. The liquid front velocity reduces irrespective of the existence of particles after the maximum velocity of the liquid front. The velocity of the liquid in case of the non-existence of particles diminishes more quickly than that in case of the existence of particles.

In contrast to the liquid front velocity in the liquid-gas-particle flows, those of the particle front are affected by PFC and PS. The profile of the particle front velocity can be divided into two groups. The velocity profile of the first group is composed of two sections. The first section has increasing velocity, while the second section has almost constant velocity. The velocity profile of the second group has increasing behavior, and reaches to the maximum. The second section has decreasing velocity.

Figure 7(b) shows the acceleration which correspond to the velocities in Fig. 7(a). The largest PS with the smallest PFC results in the largest inertia. The deceleration for the decreasing rate of the velocity is the smallest among the cases of PFC and PS considered in this study. As a result, for a smaller PFC, the inertia of the particles is preserved further downstream until the vertical wall, which leads to the increasing and constant behavior of the velocity. Otherwise, in case of the largest PFC, the magnitude of the deceleration becomes quickly increasing, which could be contributed to a larger friction force.

To cover all ranges of PFC and PS considered in this study, isometric views of the front positions of the liquid and particles are presented in Fig. 8. The coordinates for the liquid and the particles consist of (dp, μs, xl*) and (dp, μs, xp*), respectively. The liquid front position shows a plane surface at each time, as shown in Fig. 8(a), which means that the liquid front is independent of PFC and PS. However, the effect of PFC and PS on the particle front position is apparent, as shown in Fig. 8(b). The difference in the particle front position for different parameters is minor at t*=1, and the surface is nearly planar. However, as time goes by, the effect of PFC and PS on the particle front position becomes bigger, resulting in the augmentation of the difference of each particle front position according to PFC and PS. The maximum value appears at a PS of dp=0.5do and PFC of μs=0.1 at every time, which means that the particle front under these conditions has the fastest velocity among properties considered in this study. The early plane shape become a 3-D rhombus, which means that the discrepancy of the upper vertex (the maximum value) and the lower vertex (the minimum value) becomes bigger as time goes by.

FIG. 8.

Three-dimensional plot of (a) liquid and (b) particle front position at t*=1, 2, 3, and 3.5 as a function of the particle diameter ratio (dp/do) and particle friction coefficient ( μs).

FIG. 8.

Three-dimensional plot of (a) liquid and (b) particle front position at t*=1, 2, 3, and 3.5 as a function of the particle diameter ratio (dp/do) and particle friction coefficient ( μs).

Close modal

The difference between the particle front position and the liquid front position is plotted as 3-D as function of PFC and PS in Fig. 9. The dashed line indicates that the front positions of the liquid and particles are the same. The figures correspond to t*= 1, 2, 3, and 3.5. In the early stage, except for PS of dp=0.5do and regardless of PFC, the difference of the front position of the liquid and particles is close to 0% for different PFCs and PSs at t*=1.

FIG. 9.

Three-dimensional plot of the difference of the particle front and liquid front as a function of particle diameter ratio (dp/do) and particle friction coefficient ( μs) at (a) t*=1, (b) t*=2, (c) t*=3 and (d) t*=3.5. The dashed lines indicate values of 0%.

FIG. 9.

Three-dimensional plot of the difference of the particle front and liquid front as a function of particle diameter ratio (dp/do) and particle friction coefficient ( μs) at (a) t*=1, (b) t*=2, (c) t*=3 and (d) t*=3.5. The dashed lines indicate values of 0%.

Close modal

The particle front was ahead of the liquid front for PS of dp=0.5do, regardless of PFC (Fig. 9(a)). As time goes by, for more cases considered in this study, the particle front lags behind the liquid front at t*=2 except at PS=0.5, regardless of PFC. However, at t*=3, the difference of the liquid and particle front positions is close to 0%, which means that the particle front position overlapped with the liquid front position in most cases. In addition, the difference between the maximum value at dp=0.5do and μs=0.1 and the minimum value at dp=1.0do and μs=1.0 is the biggest at t*=3.

Before the particle front collides against the other side walls, the trends lasted for t*=3.5. However, the degree of difference between the maximum and minimum values is somewhat alleviated. In addition, when PFC is 0.5 and the range of PS is 0.1–0.3, the particle front is always ahead of the liquid front, while the particle front always lags behind the liquid front when the ranges of PFC and PS are 0.6–1.0 and 0.3–0.9, respectively.

The liquid has an effect of the reflection force against the approaching wall, leading to the decrease of the velocity. However, the particles keep the inertia close to the wall. Thus, the smaller particles with smaller friction coefficient keep maintains the velocity, which support the tendency of increasing and constant velocity, as discussed in Fig. 6.

In order to identify the material that arrives first on the opposite wall, the formation of the free-surface and particle distribution for representative cases are presented in Fig. 10 when particle or liquid arrive at opposite wall. To cover all ranges of PFC and PS, the arrival times of the liquid and particle fronts are quantitatively presented in Fig. 11.

FIG. 10.

Top view of free-surface and particle formation for μs=0.1 (1st column), 0.4 (2nd column) and 0.9 (3rd column) at (a) dp=0.5do, (b) dp=0.6do, (c) dp=0.7do, (d) dp=0.8do, (e) dp=0.9do and (f) dp=1.0do.

FIG. 10.

Top view of free-surface and particle formation for μs=0.1 (1st column), 0.4 (2nd column) and 0.9 (3rd column) at (a) dp=0.5do, (b) dp=0.6do, (c) dp=0.7do, (d) dp=0.8do, (e) dp=0.9do and (f) dp=1.0do.

Close modal
FIG. 11.

Arrival time of particle and liquid fronts (ta*) at the opposite wall versus particle friction coefficient (μs).

FIG. 11.

Arrival time of particle and liquid fronts (ta*) at the opposite wall versus particle friction coefficient (μs).

Close modal

At the lowest particle size of dp=0.5do among the particle sizes considered in this study, the particle front arrives earlier than the liquid front head, which is independent of the particle friction coefficient, as shown in Fig. 10(a). This early arrival of the particle fronts at the lowest particle size of dp=0.5do can be clarified by Fig. 11 where the arrival time of the liquid front is about ta*=4.0, regardless of the parameters of the particles.

Like the lowest particle size of dp=0.5do, the smallest friction coefficient of for μs=0.1 leads to earlier arrival of the particle front than the liquid front head, regardless of the particles size, as shown in the first column of Fig. 10(a-f). Thus, Fig. 11 shows smaller arrival time of the particles front for μs=0.1 than the liquid front head.

In general, at a fixed PS, The arrival time of the particle front increases with increasing PFC. However, as PS increases, the increasing rate of arrival time of the particle front with increasing PFC becomes larger. When PS was 0.5, the difference in the maximum and minimum arrival times depending on PFC WAS about 3%. However, when PS was 1.0, the difference was about 15%. This dependence of the arrival time of the particle front on PFC at a fixed PS is the same as the variation of PS at a fixed PFC. Consequently, a bigger particle size corresponded with a bigger effect of PFC on the arrival time.

Finally, we classify the particle size and the particle friction coefficient into two groups where the particle front arrives faster or not. One is liquid front first arrival and another is liquid front first arrival. These two groups are mapped in Fig. 12 where the x- and y-axes represent PFC and PS, respectively.

FIG. 12.

Classification of three groups as a function of particle diameter ratio (dp/do) and particle friction coefficient (μs).

FIG. 12.

Classification of three groups as a function of particle diameter ratio (dp/do) and particle friction coefficient (μs).

Close modal

The present study numerically investigated the liquid-gas-particles mixture flow during dam break using the CFD-DEM numerical method. The adopted numerical methods were validated through comparison with previous researches of computation and experiments. The main purpose is to investigate the effects of particle size (PS) and the particle friction coefficient (PFC) on the behavior of the particles and the free surface. To identify the effects of PFC and PS, the motion of the particles and free surface, the front position of the liquid and particles, and the arrival times were analyzed and compared for different values of PFC and PS.

The liquid front was almost independent of PFC and PS, but the particle front was strongly affected by them. The particles moved faster when simultaneously decreasing the size and friction coefficient. Thus, the smallest values of PS and PFC lead to the earliest arrival of the particle front at the opposite wall. Recently, authors’ previous study33 considered the effect of the particle density on the multiphase liquid-gas-particle flow from a dam break using CFD-DEM. We showed almost the invariance of the liquid front position and the formation of the free-surface according to the particle density. However, the particle front positions strongly depended on the particle density. The PFC and PS also gives the same effect on the liquid front position, the formation of the free-surface, and the particle front position as the particles density.

The liquid front velocity reduces irrespective of the existence of particles after the maximum velocity of the liquid front. The velocity of the liquid in case of the non-existence of particles diminishes more quickly than that in case of the existence of particles. Especially, the liquid front velocity was independent of the PS and PFC, which is consistent with the time variation of the liquid front head. The particle front velocity showed two pattern according to time. First pattern is the increasing and then saturated. Second pattern has increasing and decreasing behavior. These two patterns are mainly governed by the PFC. The smaller PFC leads to first pattern, since the inertia overcomes the weak friction, regardless of the PS. The larger PFC leads to second pattern, since the inertia becomes weak due to relatively the strong friction, regardless of the PS.

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIP) through GCRC-SOP (No. 2011- 0030662).

1.
R.
Iverson
, “
The physics of debris flows
,”
Rev Geoph.
35
,
245
296
, https://doi.org/10.1029/97rg00426 (
1997
).
2.
L.
Highland
,
S.
Ellen
,
S.
Christian
, and
W.
Brown
 III
, “
Debris-flow hazards in the United States
,” U.S. Geological Survey Fact Sheet 176-197 (
1997
).
3.
R.
Griffiths
, “
The dynamics of lava flows
,”
Annu Rev Fluid Mech
32
,
477
518
(
2000
).
4.
O.
Roche
,
S.
Montserrat
,
Y.
Niño
, and
A.
Tamburrino
, “
Experimental observations of water-like behavior of initially fluidized, dam break granular flows and their relevance for the propagation of ash-rich pyroclastic flows
,”
J Geophys Res-Solid Earth
113
,
B12203
, https://doi.org/10.1029/2008jb005664 (
2008
).
5.
C.
Ancey
, “
Plasticity and geophysical flow: A review
,”
J Non-Newton Fluid
142
,
4
35
(
2007
).
6.
N.
Balmforth
and
R.
Kerswell
, “
Granular collapse in two dimensions
,”
J Fluid Mech
538
,
399
428
(
2005
).
7.
R.
Bagnold
, “
Experiments on a gravity-free dispersion of large solid spheres in a Newtonian fluid under shear
,”
Pro R Soc London, Ser A
225
,
49
63
(
1954
).
8.
R.
Bakhtyar
,
A.
Ghaheri
,
A.
Yeganeh-Bakhtiary
, and
D.
Barry
, “
Process-based model for nearshore hydrodynamics, sediment transport and morphological evolution in the surf and swash zones
,”
Appl Ocean Res
31
,
44
56
(
2009
).
9.
H.
Wu
,
M.
Gui
,
X.
Yang
,
J.
Tu
, and
S.
Jiang
, “
A smoothed void fraction method for CFD-DEM simulation of packed pebble beds with particle thermal radiation
,”
Int J Heat Mass Tranf
118
,
275
288
(
2018
).
10.
J.
Park
,
H.
Park
,
M.
Lee
, and
K.
Moriyama
, “
Modeling of pressure drop in two-phase flow of mono-sized spherical particle beds
,”
Int J Heat Mass Tranf
127
,
986
995
(
2018
).
11.
R.
Petricek
,
T.
Moucha
,
J.
Rejl
,
L.
Valenz
,
J.
Haidl
, and
T.
Cmelikova
, “
Gas-liquid-solid volumetric mass transfer coefficient and impeller power consumptions for industrial vessel design
,”
Int J Heat Mass Tranf
121
,
653
662
(
2018
).
12.
S.
Chawdhary
,
A.
Khosronejad
,
G.
Christodoulou
, and
F.
Sotiropoulos
, “
Large eddy simulation of density current on sloping beds
,”
Int J Heat Mass Transf
120
,
1374
1385
(
2018
).
13.
Y.
Yap
,
H.
Li
,
J.
Lou
,
L.
Pan
, and
Z.
Shang
, “
Numerical modeling of three-phase flow with phase change using the level-set method
,”
Int J Heat Mass Transf
115
,
730
740
(
2017
).
14.
H.
Tang
,
T.
Keen
, and
R.
Khanbilvardi
, “
A model-coupling framework for nearshore waves, currents, sediment transport, and seabed morphology
,”
Commun Nonlinear Sci
14
,
2935
2947
(
2009
).
15.
J.
Wang
,
H.
Tang
, and
H.
Fang
, “
A fully coupled method for simulation of wave-current-seabed systems
,”
Commun Nonlinear Sci
18
,
1694
1709
(
2013
).
16.
F.
De Rooij
, “
Sedimenting particle-laden flows in confined geometries
,” Diss.,
University of Cambridge
(
2000
).
17.
L.
Guo
and
K.
Morita
, “
Numerical simulation of 3D sloshing in a liquid–solid mixture using particle methods
,”
Int J Numer Meth Eng
95
,
771
790
(
2013
).
18.
R.
Delannay
,
A.
Valance
,
A.
Mangeney
,
O.
Roche
, and
P.
Richard
, “
Granular and particle-laden flows: From laboratory experiments to filed observations
,”
J Phys D Appl Phys
50
,
053001
(
2017
).
19.
J.
Xia
,
B.
Lin
,
R.
Falconer
, and
G.
Wang
, “
Modelling dam-break flows over mobile beds using a 2D coupled approach
,”
Adv Water Resour
33
,
171
183
(
2010
).
20.
Q.
Ran
,
J.
Tong
,
S.
Shao
,
X.
Fu
, and
Y.
Xu
, “
Incompressible SPH scour model for movable bed dam break flows
,”
Adv Water Resour
82
,
39
50
(
2015
).
21.
S.
Lin
and
Y.
Chen
, “
A pressure correction-volume of fluid method for simulations of fluid-particle interaction and impact problems
,”
Int J Multiphas Flow
49
,
31
48
(
2013
).
22.
P.
Cundall
and
O.
Strack
, “
A discrete numerical model for granular assemblies
,”
Geotechnique
29
,
47
65
(
1979
).
23.
A.
Stevens
and
C.
Hrenya
, “
Comparison of soft-sphere models to measurements of collision properties during normal impacts
,”
Powder Technol
154
,
99
109
(
2005
).
24.
Y.
Tsuji
,
T.
Kawaguchi
, and
T.
Tanaka
, “
Discrete particle simulation of two-dimensional fluidized bed
,”
Powder Technol
77
,
79
87
(
1993
).
25.
X.
Sun
and
M.
Sakai
, “
Three-dimensional simulation of gas–solid–liquid flows using the DEM–VOF method
,”
Chem Eng Sci
134
,
531
548
(
2015
).
26.
L.
Guo
,
K.
Morita
, and
Y.
Tobita
, “
Numerical simulations of gas-liquid-particle three-phase flows using a hybrid method
,”
J Nucl Sci Technol
53
,
271
280
(
2016
).
27.
L.
Lucy
, “
A numerical approach to the testing of the fission hypothesis
,”
Astron J
82
,
1013
1024
(
1977
).
28.
Y.
Yamada
,
M.
Sakai
,
S.
Mizutani
,
S.
Koshizuka
,
M.
Oochi
, and
K.
Murozono
, “
Numerical simulation of three-dimensional free-surface flows with explicit moving particle simulation method
,”
Trans Atom Energ Soc Jpn
10
,
185
193
(
2011
).
29.
X.
Sun
,
M.
Sakai
, and
Y.
Yamada
, “
Three-dimensional simulation of a solid–liquid flow by the DEM–SPH method
,”
J Comput Phys
248
,
147
176
(
2013
).
30.
OpenCFD Ltd. OpenFOAM - The Open Source CFD Toolbox (2009) URL: http://www.openfoam.com.
31.
C.
Kloss
,
C.
Goniva
,
A.
Hager
,
S.
Amberger
, and
S.
Pirker
, “
Models, algorithms and validation for opensource DEM and CFD–DEM
,”
Progress in Computational Fluid Dynamics, an International Journal
12
,
140
152
(
2012
).
32.
C.
Goniva
,
K.
Gruber
, and
C.
Kloss
, “
Sediment erosion a numerical and experimental study
,” River Flow
415
422
(
2012
).
33.
K.
Park
,
H.
Yoon
, and
M.
Kim
, “
CFD-DEM based numerical simulation of liquid-gas-particle mixture flow in dam break
,”
Commun Nonliear Sci
59
,
105
121
(
2018
).
34.
C.
Gougon
,
B.
Dalloz-Dubrujeaud
, and
N.
Thomas
, “
Bidisperse granular avalanches on inclined planes: A rich variety of behaviors
,”
The European Physical Journal E.
23
,
199
(
2007
).
35.
C.
Goniva
,
C.
Kloss
,
A.
Hager
, and
S.
Pirke
, “
An open source CFD-DEM perspective
,” In
Proceedings of OpenFOAM Workshop
22
24
(
2010
).
36.
R.
Issa
, “
Solution of the implicitly discretized fluid flow equations by operator-splitting
,”
J Comput Phys
42
,
40
65
(
1986
).
37.
I.
Janosi
,
D.
Jan
,
K.
Szabo
, and
T.
Tel
, “
Turbulent drag reduction in dam-break flows
,”
Exp Fluids
37
,
219
229
(
2004
).