numerical simulation of the in-line pressure jig unit in coal preparation minerals engineering

Upload: nhfreire

Post on 05-Apr-2018

221 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    1/12

    Numerical simulation of the in-line pressure jig unit in coal preparation

    K.J. Dong a, S.B. Kuang a, A. Vince b, T. Hughes c, A.B. Yu a,*

    a Lab. for Computer Simulation and Modelling of Particulate Systems, University of New South Wales, NSW, Australiab Elsa Consulting Group Pty Ltd., Queensland, Australiac Gekko Systems, Victoria, Australia

    a r t i c l e i n f o

    Article history:Received 18 August 2009

    Accepted 21 October 2009

    Keywords:

    Gravity concentration

    Classification

    Coal

    Computational fluid dynamics

    Discrete element method

    a b s t r a c t

    This paper presents a numerical study of the multiphase flow in an in-line pressure jig (IPJ), which is ahigh yield and high recovery gravity separation device widely used in ore processing but may have poten-

    tial in coal preparation. The mathematical model is developed by use of the combined approach of com-

    putational fluid dynamics (CFD) for liquid flow and discrete element method (DEM) for particle flow. It is

    qualitatively verified by comparing the calculated and measured results under similar conditions. The

    effects of a few key variables, such as vibration frequency and amplitude, and the size and density of rag-

    ging particles, on the flow and separation performance of the IPJ are studied by conducting a series of

    simulations. The results are analyzed in terms of velocity field, porosity distribution and forces on parti-

    cles. The findings would be helpful in the design, control and optimisation of an IPJ unit.

    2009 Elsevier Ltd. All rights reserved.

    1. Introduction

    The use of jigging machinery for the classification and benefici-

    ation of ore has a long history. Classic jigging units characteristi-

    cally dilate the particle bed by an upward blast of water caused

    by the movement of a remote piston through a screen. Particles

    of different densities are then likely to segregate when they settle.

    Repeating such an operation makes the lighter particles remain on

    the top layer and the heavier particles drop down to the bottom

    layer. These particles can then be collected at either end to meet

    specific product requirements. These units were popular during

    and prior to the 1980s. In the 1990s, the jigging unit was improved

    by incorporating a centrifugal action in the unit (Beniuk et al.,

    1994). However, recent technological developments have resulted

    in jigging technology becoming an even more sophisticated tool of

    classification. For example, the invention of the in-line pressure

    jig (IPJ) resulted in a more sophisticated classifier and can achieve

    even higher levels of efficiency. When using this method, a screen

    is moved up and down in a cyclic manner by means of a hydrauli-

    cally powered servo that is mechanically linked to the screen.

    Moreover, the entire process occurring in a confined pressurized

    environment, adding a new dimension of security to the unit.

    During the last decade, the IPJ has grown extensively in its tech-

    nology in applications in the metalliferous industry. More recently,

    it is being considered as an alternative means to the dense medium

    cyclone for processing coal particles in large size ranges (0.25

    30 mm). Some pilot scale tests have been performed to investigate

    the effects of the operational conditions for optimization of the

    control of IPJ in such separations (Vince et al., 2007). However,

    due to the complicated nature of the system and the number of

    the parameters involved, the full optimization through experimen-

    tal studies is not an easy task. The lack of the fundamental under-

    standings of such processes is the key motivation for a theoretical

    study.

    There are few fundamental studies on the classification mecha-

    nism of the jigging devices in the current literature. Steiner (1996)

    studied the classical jigging device with only bare basics being de-

    bated. Galvin et al. (2002) and Mishra and Adhikari (1999) investi-

    gated the water flow in the jigging process in a simple geometry.

    Nesbitt et al. (2005) discussed only the effects of vibrating condi-

    tions on the jigging process in IPJ, although other parameters such

    as the properties of the ragging particles on the screen are also very

    critical.

    In principle, the bulk behavior of particles in a system depends

    on the collective outcome of the interactions between individual

    particles, particles and boundary walls, and particles and fluid.

    Therefore, an investigation of the particle flow inside an IPJ on a

    particle scale should provide insight into the classification mecha-

    nism of the unit. Experimentally, such an investigation is challeng-

    ing because the access to an IPJ is difficult being a confined

    pressurized unit. However, numerical simulation based on the

    so-called discrete element method (DEM) (Cundall and Strack,

    1979) provides an effective away to perform such studies. This

    method has been applied in the study of particlefluid flow pro-

    cesses in various industrial processes and is shown to be very

    0892-6875/$ - see front matter 2009 Elsevier Ltd. All rights reserved.doi:10.1016/j.mineng.2009.10.009

    * Corresponding author. Tel.: +61 2 93854429; fax: +61 2 93855956.

    E-mail address: [email protected] (A.B. Yu).

    Minerals Engineering 23 (2010) 301312

    Contents lists available at ScienceDirect

    Minerals Engineering

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / m i n e n g

    http://dx.doi.org/10.1016/j.mineng.2009.10.009mailto:[email protected]://www.sciencedirect.com/science/journal/08926875http://www.elsevier.com/locate/minenghttp://www.elsevier.com/locate/minenghttp://www.sciencedirect.com/science/journal/08926875mailto:[email protected]://dx.doi.org/10.1016/j.mineng.2009.10.009
  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    2/12

    useful in understanding the fundamentals (Zhu et al., 2007, 2008).

    In particular, it has been adapted for modeling the vibrating

    screening process (Dong et al., 2009).

    In this work we present a three-dimensional CFDDEM model,

    which is capable of simulating an IPJ unit. The model is validated

    by comparing the calculated and measured results under similar

    conditions. The effects of a few key variables on the flow and sep-

    aration performance of the IPJ are studied by conducting a series of

    controlled numerical experiments, the key variables being the

    vibration conditions and properties of ragging particles. The

    numerical results are analyzed in terms of forces on particles, par-

    ticle and fluid velocities and porosities of the particle bed, which

    present a better understanding on the particlefluid flow in the

    IPJ unit.

    2. Model description

    Fig. 1a shows the working principle of the IPJ schematically. For

    confidential reasons, the detailed dimensions are not given here.

    The whole unit is sealed, hutch water and slurry (including water

    and coal particles) is pumped in. Ragging particles are put onto

    the screen. The upper part of the IPJ, including the upper part of

    the inner chamber with ring shape apertures on the wall, the

    screen and the feeding bowl, is continuously vibrated with jig-

    saw motions. Coal particles are fed from the top tube into the IPJ,

    and they either flow out through the apertures on the inner wall

    and then to the product outlet, or pass through the screen and dis-

    charged from the tails or reject outlet.

    A coupled CFDDEM model is developed here to model the sys-

    tem. In DEM, the particle flow is treated as a discrete phase, and

    the translational and rotational motions of particles are deter-

    mined by Newtons law of motion, which can be written as

    midvidt

    fpf;i Xki

    j1

    fc;ij fd;ij mig 1

    and

    Iidxidt

    Xki

    j1

    Tij 2

    where mi, Ii, ki, vi, and xi are, respectively, the mass, momentum of

    rotational inertia, number of contacting particles, translational and

    rotational velocities of particle i; ffp,i and migare the force between

    particle and fluid and gravitational force, respectively; and fc,ij and

    fd,ij, and Ti,j are the contact force, viscous contact damping forceand torque between particles i and j. These individual interaction

    forces and torques are summed over the ki particles in interaction

    with particle i. The particleparticle or particlewall contact force

    is calculated according to non-linear models commonly used in

    DEM, as recently reviewed by Zhu et al. (2007). The particlefluid

    interactions include the buoyancy force and the drag force. The drag

    force is calculated according to Di Felices correlation (1994). The

    equations used to calculate the forces and torques involved in

    Eqs. (1) and (2) can be found elsewhere (Dong et al., 2008; Kuang

    et al., 2008).

    In CFD, the water flow is treated as a continuous phase and

    modeled in a way similar to the one in the conventional two-fluid

    modeling. Thus, its governing equations are the conservation of

    mass and momentum in terms of local mean variables over a com-putational cell, given by

    r qfu 0 3

    and

    r qfuu rP r s qfg Fpf 4

    where q, u, P and Fpf are, respectively, the fluid density, velocity,

    pressure, and the volumetric forces between particle and fluid; s

    is fluid viscous stress tensor, calculated according to standard ke

    turbulent model.

    DEM is solved by an object-oriented-programming based in-

    house code which can handle dynamic and complex boundaries

    and calculate the fluidparticle forces with the fluid flow field

    introduced from CFD simulation (Dong et al., 2008). The model

    has been successfully used in the simulation studies of complicated

    (a) (b)

    Fig. 1. (a) Schematic representation of in-line pressure jig and (b) the mesh used in CFD.

    302 K.J. Dong et al./ Minerals Engineering 23 (2010) 301312

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    3/12

    screening processes (Dong et al., 2009). CFD is solved by the com-

    mercial package fluent. The IPJ unit contains a big number of par-

    ticles and CFD meshes and generally takes about 30 s to achieve

    steady-state coal separation. Because of computational limit at this

    stage of development, only 1/8th sector of the region around the

    actual jigging area of the in-line pressure jig is considered in the

    DEM simulations as shown in Fig. 2. However, the full unit is sim-

    ulated in CFD, assuming that the water flow is a steady-state flow

    and is not affected by screen vibration. The water flow over screen

    is largely axisymmetric. To eliminate inconsistencies between CFD

    and DEM due to the use of different computational domains, fluid

    field is first averaged over all the sectors before being used to cal-

    culate particlefluid forces. In addition, porous medium instead of

    two-way coupling of CFD and DEM is used here to account for ef-

    fects of particles on the fluid flow. In this work, porous medium is

    only applied to the region over the screen as shown in Fig. 3, where

    the solid loading is relatively high and its water flow is crucial to

    the coal separation. The porous medium considers the presence

    of particles in fluid by the addition of volumetric particlefluid

    forces to the standard momentum equations of fluid. Here, the vol-

    umetric forces are estimated according to Ergun equation, which

    are composed of two parts: a viscous loss term and an inertial loss

    term:

    Fpf lf

    au C

    1

    2qfjuju 5

    where a d2p

    150

    e3f

    1ef2; C

    3:5dp

    1ef

    e3f

    ; ef is porosity, which is set to 0.45

    in Zone I and 0.7 in Zone II according to the particle configuration, in

    a preliminarily numerical experiment; and dp is the maximum par-

    ticle size in each zone.

    3. Simulation conditions

    In this work, the simulation is based on a laboratory-scale IPJ

    unit and the conditions used in the related work (Vince et al.,

    2007). The water flow rates at the inlets and outlets are specified

    according to the experimental measurements. The actual size dis-

    tribution of coal particles (from 2 to 6 mm) is simplified to 3 differ-

    ent sized particles (2, 4 and 6 mm) and their density distribution is

    assumed to be uniformly distributed from 1.2 RD to 1.9 RD. Note

    that relative density (RD) is used here as the density unit in this

    work asit is often used intheIPJ studies(Nesbitt et al., 2005; Vince

    et al., 2007). A summary of the conditions used in the simulation is

    listed in Table 1. The variables have their base values correspond-

    ing to the experiment conditions. The vibration frequency and

    amplitude and ragging density and size are varied as shown in

    the range column when studying their effects on the separationperformance. When one variable is changed, other variables are

    all kept to their base values. Table 2 lists the parameters used in

    the DEM simulation, which are generally based on the properties

    of the coal.

    When a simulation run begins, the IPJ is full of water, and the

    screen is kept stationary. The ragging particles are first placed on

    the screen. The screen vibration is then switched on, and the coal

    particles begin to be generated simultaneously in the top part of

    the feed tube. Coal particles then flow in the IPJ unit, and those that

    Fig. 2. Schematic representation of the region (under shadows) considered in the DEM simulation: (a) top view and (b) side view.

    Zone I

    Zone II

    Zone I

    Zone II

    Fig. 3. Illustration of porous media zones.

    Table 1

    List of conditions used in this work.

    Variables (Unit) Base Value Range

    Hutch Water(l/s) 5 -

    Reject Flow (l/s) 3.3

    Vibration frequency (Hz) 2 1, 2, 4

    Vibration Amplitude (mm) 10 5, 10, 20

    Solid Feed Rate (tph) 1.0 -

    Size of Coal Particles (mm),

    and the volume fraction

    2.0, 20% -

    4.0, 30%

    6.0, 50%

    Density of Coal Particles (RD),

    and the volume fraction

    1.2, 12.5% -

    1.3, 12.5%

    1.4, 12.5%

    1.5, 12.5%

    1.6, 12.5%

    1.7, 12.5%

    1.8, 12.5%

    1.9, 12.5%

    Ragging Density (RD) 1.6 1.4, 1.6, 1.8

    Ragging Diameter (mm) 18 16, 18, 20

    K.J. Dong et al./ Minerals Engineering 23 (2010) 301312 303

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    4/12

    exit through the apertures on the inner wall are collected in the

    product flow, while those that exit from the bottom of the simu-

    lated zone are collected in the reject flow. A simulation run needs

    to be carried for a certain time to achieve the steady state, which

    means the total outflow rate (the product flow plus the reject flow)

    is equal to the feed rate for all kinds of particles. The data used in

    the analyses are obtained after 30 s simulations. For most cases,

    the steady state is achieved at this stage. However for several cases

    the steady state is not even achievable after 50 s. These cases are

    not continued, while the data obtained from them is used in a lim-

    ited range in the latter discussions.

    4. Results and discussion

    4.1. Water flow

    Firstly, we discuss the water flow field modeled by CFD. Fig. 4

    shows the representative flow field in the IPJ, along with details

    around the feed bowl, the region considered in DEM. The figure

    shows that the water velocity in the unit overall is small except in-

    side the feed bowl and near the bottom area. The trajectory of

    water exiting the feed well covers a significant part of the screen

    surface. This flow would tend to push coal particles away from

    the region close to the inner surface of the screen towards the out-

    er surface. In addition, a significant recirculation zone on the main

    chamber is found, which is believed to result from the strong hor-izontal flow of hutch water. This recirculation zone may hinder the

    settling down process of tails in the inner chamber, and is thus not

    favored in an IPJ.

    In order to further understand the flow in the IPJ, we trace the

    flow paths of water using massless particles separately. The results

    are shown in Fig. 5. It can be seen from Fig. 5a that the feeding

    water always exits via the products outlet, and none of it is found

    to exit from the tails outlet. On the other hand, the pathways fol-

    lowed by hutch water as shown in Fig. 5b show that most of the

    hutch water report to tailings with a small proportion passing

    through the particle bed and over the screen and discharge with

    the product. Some hutch water also circulates in the unit before

    exiting via the tails and product outlets. Evidently, the strong hor-

    izontal flow of hutch water helps push tails towards the tails out-

    let, and its portion flowing over the screen may raise coal particles

    towards the screen surface, and thus improving separation perfor-

    mance. Note that the hutch water also results in harmful circulat-

    ing flows, as discussed above. This two-sided role of hutch water

    explains why, in the previous experiments (Vince et al., 2007),

    the increase of hutch water cannot always increase the separation

    efficiency of the unit.

    Fig. 6 shows the flow fields of water in the feed bowls with dif-

    ferent depths. Even with the effects of particles ignored, it can be

    seen from the figure that water flow is very sensitive to the geom-

    etry of feed bowl, which is used to distribute slurry into the screen.

    The water flow in the unit with a deep bowl has a greater circulat-

    ing flow, which may trap particles and cause difficulty for coal sep-

    aration. This effect has not been observed in the experiments, and

    deserves further investigation in the future.

    In the following particle flow simulations, to simplify the stud-

    ies of the first stage work, we focus our studies on the cases using

    the water flow field modeled with base conditions; although other

    cases are also studied, they will be reported elsewhere.

    4.2. Particle flow

    To validate the numerical model, the simulation results of the

    base case are compared with those measured in terms of partition

    number of particles in the reject flow as a function of particle den-

    sity; partition number is defined as the ratio of the mass of parti-

    cles found in the reject flow to the mass of the fed particles for

    particles of a certain density. The results are shown in Fig. 7a.The predicted results are qualitatively comparable to the experi-

    mental results. In addition, in the following parametric studies,

    changes of the product collection rate with the changes of each

    variable are also found to be qualitatively comparable with the

    experimental findings. These agreements demonstrate the validity

    of our numerical model, although no effort has been made to tune

    Table 2

    List of parameters used in DEM simulations.

    Youngs modulus of particles (N/m2) 1 107

    Youngs modulus of walls (N/m2) 1 107

    Damping coefficient (inter-particle) 1 108

    Damping coefficient (particlewall) 2 105

    Sliding friction coefficient 0.4

    Rolling friction coefficient 0.01

    Fig. 4. Flow field obtained by CFDcalculation on theyzplane atx = 0: (a)spatial distribution of water velocity, (b) streamline with velocity contours, and (c)velocity field inthe amplified section in (b), when hutch water flowrate = 5 l/s, reject flowrate = 3.3 l/s, and slurry flowrate = 4.61 l/s.

    304 K.J. Dong et al./ Minerals Engineering 23 (2010) 301312

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    5/12

    parameters for better quantitative agreement. Moreover, Fig. 7b

    shows that the partition curves for different sized particles are

    similar to that of the sum of all sized particles, indicating that

    the IPJ is suitable for different sized particles as well, although itcan also be noted that the increase of the particle size decreases

    the cut point and sharpens the partition curve.

    Before the parametric studies by a series of controlled numeri-

    cal simulations, we first focus our discussion on the base case.

    Fig. 8 shows the snapshots of the flow of coal particles in the IPJ.

    The particle bed on the screen is a mixture of ragging particles

    and coal particles of different densities and sizes. It is thought that

    the particle bed is critical to the segregation of particles and the

    subsequent separations (Vince et al., 2007; Nesbitt et al., 2005).

    We therefore investigate the bed structure in terms of the time

    averaged spatial distributions of volume fractions of particles for

    different kinds of particles. The results are shown in Fig. 9. Since

    the particle flow is largely axisymmetric, we show the averaged re-

    sults in the radial and axial directions (donoted as r and z,respectively).

    It can be seen from Fig. 9 that in the particle bed, particles of dif-

    ferent sizes have similar distributions, while particles of different

    densities have different distributions. The lightest particles mostly

    distribute away from the center and stick to the inner wall. In par-ticular, those at the top are close to the apertures on the inner wall,

    therefore they are easy to report to the product flow through these

    apertures. Contrary to this, the heaviest particles are more likely to

    accumulate close to the center, close to the feeding bowl, and in a

    much lower position, away from the apertures on the inner walls.

    The medium dense particles distribute more sparsely in all regions.

    The segregation among particles of different densities but not of

    different sizes could be related to the similar partition curves of

    different sized particles in Fig. 7b. This further indicates the appli-

    cability of the IPJ to separate particles according to their densities.

    The segregation of the particles with different densities should

    occur because of their different motions in the IPJ. From Fig. 8a, we

    can see general differences of their trajectories. When particles are

    fed into the feeding bowl, they follow similar paths. However, thedifference emerges when particles are bumped out from the feed-

    Fig. 5. Flow paths of water entering at the feed (a) and hutch water inlets (b), corresponding to Fig. 4.

    Fig. 6. Effect of feed bowl geometry on water flow without considering particles: (a) real feed bowl geometry and (b) modified feed bowl geometry.

    K.J. Dong et al./ Minerals Engineering 23 (2010) 301312 305

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    6/12

    ing bowl by the vibration and water flow. It can be seen that the

    trajectories of the lighter particles are higher than those of the hea-

    vier particles. Since the water flow on the top part of the IPJ show

    high outwards radial velocity in Fig. 4c, some of the lighter parti-

    cles are likely to be drawn directly to the product flow through

    the apertures on the inner wall and by-pass the particle bed on

    the screen. These particles appear to be separated by an elutria-tion-based mechanism. Other particles, however, fall on the parti-

    cle bed first. And then with the jigging of the particle bed, they

    either flow up and are drawn to the product flow through the aper-

    tures, or percolate through the jigging bed and report to the reject

    flow. Hence, in the simulated system, there are two mechanisms

    for the separation of lighter particles to the product flow, corre-

    sponding to different flow paths.

    In order to see the behavior of different particles more clearly in

    simulations, we traced a group of particles which were generated

    at the same time. From the tracing of this group of particles, four

    particles with typical trajectories are chosen for discussion. These

    are indexed as particle A, B, C and D, respectively. The sizes of all

    of them are 4 mm while their densities are 1.2 RD, 1.5 RD, 1.6 RD

    and 1.8 RD, respectively. The same sized particles have been cho-sen because different sized particles give similar relationships be-

    tween the partition number and density as seen from Fig. 7b, and

    the partition numbers of 4 mm particles are closest to the summed

    partition number. In the following analysis, we consider only the

    movements of particles along the radial and axial directions, as

    those along the tangential direction are negligible.

    Fig. 10a shows the entire trajectory of particle A in IPJ, and the

    trajectories of particles B and D before they fall onto the jiggingbed. The trajectory of particle C is not shown here for it is very sim-

    ilar to that of particle B, as their densities are close. From the figure

    we can see that particle A directly reports to the product flow

    while the other two particles (B and D) fall down onto the particle

    bed first. In this stage, particles have few collisions with other par-

    ticles; therefore liquidparticle interactions should dominate the

    motion of the particle. From Fig. 10c wecan see that when particles

    are bumped out from the feeding bowl, the axial liquidparticle

    forces for particle A are upwards (indicating positive value), while

    those for particles B and C are downwards (indicating minus value)

    or slightly upwards. Fig. 4c shows that the fluid velocities in this

    region flow mainly along the radial direction, so the drag force

    on the particles in the axial direction should be rather small.

    Hence, the buoyancy force should be the main axial force on theparticles. As the buoyancy force is relatively larger for lighter par-

    (a) (b)

    Fig. 7. Partition number of particles in the reject flow as a function of density: (a) comparison between simulation and experimental results (from Vince et al., 2007) and (b)

    numerical results for different sized particles and the sum of all sized particles.

    Fig. 8. Snapshots of particles in IPJ when screen is at: (a) the highest position and (b) the lowest position. The rectangle in (a) indicates the region investigated in Fig. 9.

    306 K.J. Dong et al./ Minerals Engineering 23 (2010) 301312

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    7/12

    ticles, particle A obtains a relatively large upwards force to support

    its gravity and has the chance to flow directly out of the inner

    chamber.

    After particles fall onto the particle bed, the situation becomesmore complicated. From Fig. 11 we can see that although at the

    end, particle B reports to the product flow, while particles C and

    D flow down towards the reject flow, their trajectories in the pro-

    cess are rather random, especially for particles with medium den-

    sities (B and C). Their velocities and forces also show very

    disordered patterns with random variations of the direction along

    both radial and axial directions in Fig. 12. Nevertheless, the main

    difference in the movement of these particles is along the axial

    direction, since it determines whether they report to the productflow or to the reject flow. Thus we mainly focus our analysis on

    the axial direction. The right column of Fig. 12 shows the compar-

    isons between the particlefluid forces, particleparticle forces and

    total forces on a particle along the axial direction. We can see that

    the total axial forces oscillate about a value a bit lower than zero

    250

    270

    290

    310

    330

    350

    370

    390

    410

    430

    0 50 100 150 200 250

    Z(mm)

    r (mm)

    Particle A

    Particle B

    Particle D-1.0

    -0.8

    -0.6

    -0.4

    -0.2

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    20.2 20.4 20.6 20.8 21.0 21.2 21.4

    Force(mg)

    Time (sec)

    ParticleA

    Particle B

    Particle D

    (a) (b)

    Fig. 10. (a) Trajectories and (b) particleliquid forces along axial direction for particle A, B and D.

    Fig. 9. Time averaged space distributions of the volume fraction of different kinds of particles: (a) d = 2 mm, (b) d = 4 mm, (c) d = 6 mm, (d)q = 1.2 RD, (e) q = 1.5 RD, and (f)

    q = 1.9 RD. Coordinates are transferred to radial dimension (r) and axial dimension (z).

    K.J. Dong et al./ Minerals Engineering 23 (2010) 301312 307

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    8/12

    and particlefluid axial forces oscillate a bit lower than 1 mg, i.e.

    particle weight. The axial forces due to particleparticle collision,

    on the other hand, only show some sharp peaks either upwards

    or downwards. Briefly speaking, the liquidparticle interactions

    seem to continuously counteract gravity, while the collision forces

    make the total forces more disperse.

    A clearer picture about the forces can be seen from Fig. 13, in

    which we plot the distributions of the total force and particlefluid

    force in the axial direction for the three particles. Note, although

    there are large forces occasionally, their probabilities are rather

    small. So here we only show the force distributions in the high

    probability range. For the total force, it can be seen in Fig. 13a that

    the distributions are all bell-shaped. But the peak shifts to the right

    (at higher forces) and broadens with the decrease of particle den-

    sity (D>C>B). For the heavy particle D, the position of the peak is

    negative with a sharpshape, while the distribution of positive force

    is almost nil. This indicates that the forces on particle D in the jig-

    ging process are mainly downwards; hence it falls fastest as shown

    in Fig. 11. For particle C, the peak of the force shifts to the right

    although it is still negative, and the peak widens. A certain amount

    of forces are positive values in the distribution. Hence particle C

    has some upwards movements in the jigging process in Fig. 11.

    For particle B, the peak is at almost zero, while it can be noticed

    that the distribution of positive values cover a larger area than that

    of minus values. Thus the particle gets a better chance to drift up-

    wards and finally to be drawn to the product flow. These patterns

    are similar to the distribution of the fluidparticle forces in

    Fig. 13b, although all the curves shift to right for 1 mg. This shows

    that the fluidparticle forces and gravity are almost balanced indi-

    cating that the collision forces, acting as pulse forces, are responsi-

    ble for the jigging and the separation of coal particles in the

    particle bed.

    4.3. Effects of some operational parameters

    By means of the numerical model, we have studied the effects of

    some key operational parameters on the separation performance of

    IPJ, including the vibration conditions and properties of the ragging

    particles.

    Fig. 14 shows the partition curve for the reject flow at different

    vibration frequencies or amplitudes. We also performed the simu-

    lation when vibration amplitude is at 5 mm, while the system can-

    not achieve a steady state for a long simulation beyond 50 s due to

    the chocking of the screens. Consequently this case is not included

    here. From the figure, it can be seen that at higher frequencies or

    amplitudes, a lower cut point and a slightly sharper separation

    are achieved.

    240

    260

    280

    300

    320

    340

    360

    240

    260

    280

    300

    320

    340

    360

    240

    260

    280

    300

    320

    340

    360

    150 170 190 210 230 250

    Z(m)

    r (m)150 170 190 210 230 250

    r (m)150 170 190 210 230 250

    r (m)

    Particle B Particle C Particle D

    Fig. 11. Trajectories in jigging process for particles B, C and D. The arrows show the entrance and exit of particles.

    -0.20

    0.00

    0.20

    0.40

    0.60

    0.80

    Vr

    (m/s)

    -0.50

    -0.30

    -0.10

    0.10

    0.30

    0.50

    Vz

    (m/s)

    -5.00

    -3.00

    -1.00

    1.00

    3.00

    5.00

    20.50 21.50 22.50 23.50 24.50 25.50 26.50 27.50 28.50 20.50 21.50 22.50 23.50 24.50 25.50 26.50 27.50 28.50

    Ftotal,r

    (mg)

    Time (sec)

    -2.00

    -1.00

    0.00

    1.00

    2.00

    F

    total,z

    (mg)

    -1.00

    0.00

    1.00

    2.00

    3.00

    Fliquid,z

    (mg)

    -2.00

    -1.00

    0.00

    1.00

    2.00

    Fcontact,z

    (mg)

    Time (sec)

    Fig. 12. Velocities and forces in the radial and axial directions as a function of time for particle B.

    308 K.J. Dong et al./ Minerals Engineering 23 (2010) 301312

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    9/12

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    10/12

    sharpness of the partition curve. The case of ragging particles of

    16 mm has also been simulated, but the partition curve is not ob-

    tained as the steady state is not achievable after 50 s of simulation.

    However, it has been included in the analysis of particle velocity

    and porosity of the particle bed in Fig. 17b. It can be seen from

    Fig. 17b that larger ragging particles are more difficult to agitate

    with vibration and liquid flow, leading to a denser packing bed

    which makes coal particles harder to pass through the bed to reject

    flow. For smaller sized ragging particles (16 mm), even the velocity

    of ragging particles is greater than that of the larger ragging parti-

    cles, the particle bed still becomes denser. The reason for this is

    that the smaller the ragging particles, the smaller the gaps between

    the ragging particles, which makes coal particles harder to pass

    through the screen. Thus the screen is chocked and the steady state

    is not reached even after a long period of simulation. The compli-

    cated effects of the ragging size may need further studies as cur-

    rently there are no comparable experimental studies available.

    For all the cases studied, it is found that the sharpening of the

    partition curve is always accomplished with the decrease of the

    cut point. Fig. 18 summarizes the relationships between the cut

    point and Ep, which is the shape factor of the partition curve de-

    fined as Ep = (q75 q25)/2, where q75 and q25 are the densities

    when the partition numbers are 75%and 25% in the partition curve,

    respectively. The figure shows that Ep is strongly dependent on the

    cut point for the conditions simulated in this work. This indicates

    that to obtain good separation performance, there cannot be a high

    cut point. This relationship may need further confirmation under

    different conditions.

    (a) (b)

    Fig. 15. Average velocities of ragging particles and porosity of the center part in the particle bed at: (a) different vibration frequencies and (b) different vibration amplitudes.

    (a) (b)

    Fig. 16. Partition number as a function of particle density for the particles in the reject flow with ragging particles of: (a) different densities and (b) different sizes.

    310 K.J. Dong et al./ Minerals Engineering 23 (2010) 301312

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    11/12

    5. Conclusions

    A one-way coupled CFD and DEM model has been developed to

    simulate the flow and separation of an IPJ. The model can generate

    results qualitatively comparable to the experimental data, and

    yield microscopical information helpful to the improvement of

    the fundamental understandings and the industrial design of IPJ.

    The separation mechanisms are studied by analyzing the spatial

    distributions of particles in the IPJ and tracing typical particles in

    the process. It is found that particles of different densities distrib-

    ute differently while particles of different sizes distribute similarly

    in the IPJ, which confirms the ability of the IPJ to separate particles

    according to their densities. Two separation mechanisms are

    found. Some very light particles are drifted directly to the product

    flow mainly due to the buoyancy force and fluid drag force. On the

    other hand, most particles fall on the packed bed and are separated

    in the jigging process, in which the particlefluidforces and gravity

    are nearly balanced, while the particleparticle collision forces act

    as pulse forces and are responsible for the jigging and the

    separation.

    The effects of variables such as the vibration amplitude and fre-

    quency, and the size and density of ragging particles have been

    studied by the model. The increase of vibration frequency or ampli-

    tude introduces more mechanical energies to the particle bed, and

    the decrease of the density of ragging particles increases the effects

    of fluidparticle interactions. Consequently they all increase the

    velocities of ragging particles and the porosity of the jigging bed,

    resulting in the decrease of the cut point and the sharpening of

    the partition curve. On the other hand, the decrease of the size ofthe ragging particles leads to the increase of the velocity of ragging

    particles but also the decrease in the porosity of the jigging bed,

    which shows a more complicated effect on the partition curve.

    For all the cases studied, there is a linear correlation between the

    cut point and Ep of the partition curve.

    It is also found that the flow in IPJ is very sensitive to opera-

    tional conditions and material properties of particles, which may

    lead to unsteady state operation. There is a need to study this issue

    further.

    Acknowledgement

    The authors are grateful to Australian Coal Association ResearchProgram (ACARP) for the financial support of this work.

    References

    Beniuk, V.G., Vadeikis, C.A., Enraght-moony, J.N., 1994. Centrifugal jigging of gravity

    concentrate and tailing at Renison limited. Minerals Engineering 7, 577589.

    Cundall, P.A., Strack, O.D.L., 1979. A discrete numerical model for granular

    assemblies. Geotechniques 29, 47.

    Di Felice, R., 1994. The voidage function for fluidparticle interaction systems.

    International Journal of Multiphase Flow 20, 153.

    Dong, K.J., Guo, B.Y., Chu, K.W., Yu, A.B., Brake, I., 2008. Simulation of liquidsolid

    flow in a coal distributor. Minerals Engineering 21, 789796.

    Dong, K.J., Yu, A.B., Brake, I., 2009. DEM simulation of particle flow on a multi-deck

    banana screen. Minerals Engineering 22, 910920.

    Galvin, K.P., Pattern, S.J., Lambert, N., Callen, A.M., Lui, J., 2002. Influence of a jigging

    action on the gravity separation achieved in a teetered bed separator. MineralsEngineering 15, 11991202.

    (a) (b)

    Fig. 17. Average velocities of ragging particles and porosity of the center part in the particle bed using ragging particles of: (a) different densities and (b) different sizes.

    Fig. 18. Ep as a function of cut point.

    K.J. Dong et al./ Minerals Engineering 23 (2010) 301312 311

  • 7/31/2019 Numerical Simulation of the in-line Pressure Jig Unit in Coal Preparation Minerals Engineering

    12/12

    Kuang, S.B., Chu, K.W., Yu, A.B., Zou, Z.S., Feng, Y.Q., 2008. Computational

    investigation of horizontal slug flow in pneumatic conveying. Industrial and

    Engineering Chemistry Research 47, 470480.

    Mishra, B.K., Adhikari, B., 1999. Analysis of fluid motion during jigging. Minerals

    Engineering 12, 14691477.

    Nesbitt, A.B., Breytenbach, W., van der Plas, P.J., 2005. Characterisation of the pulse

    wave of an inline pressure jig in a near density application. Minerals

    Engineering 18, 17.

    Steiner, H.J., 1996. A contributionto thetheory of jigging, part I: similarity criteria of

    the motion of jig layers. Minerals Engineering 9, 675686.

    Vince, A., Purdon, P., Gibson, L. and Hughes, T., 2007. ACARP Project C16042

    Evaluation of in-line pressure jig, December.

    Zhu, H.P., Zhou, Z.Y., Yang, R.Y., Yu, A.B., 2007. Discrete particle simulation of

    particulate systems: theoretical developments. Chemical Engineering Science

    62, 33783396.

    Zhu, H.P., Zhou, Z.Y., Yang, R.Y., Yu, A.B., 2008. Discrete particle simulation of

    particulate systems: a review of major applications and findings. Chemical

    Engineering Science 63, 57285770.

    312 K.J. Dong et al./ Minerals Engineering 23 (2010) 301312