You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/260315478

Chute flow as a means of segregation


characterization

Article in Powder Technology · April 2014


DOI: 10.1016/j.powtec.2014.01.092

CITATIONS READS

7 53

2 authors:

Tathagata Bhattacharya Joseph Mccarthy


ArcelorMittal University of Pittsburgh
51 PUBLICATIONS 87 CITATIONS 84 PUBLICATIONS 2,036 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Tathagata Bhattacharya on 24 July 2016.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Powder Technology 256 (2014) 126–139

Contents lists available at ScienceDirect

Powder Technology
journal homepage: www.elsevier.com/locate/powtec

Chute flow as a means of segregation characterization


Tathagata Bhattacharya, J.J. McCarthy ⁎
Department of Chemical and Petroleum Engineering, University of Pittsburgh, Pittsburgh, PA 15261, USA

a r t i c l e i n f o a b s t r a c t

Article history: Handling of granular materials involves the use of many solids processing devices such as chutes, hoppers, tum-
Received 26 July 2013 blers or other transfer equipment. Unfortunately, granular materials segregate if they are subjected to flow or ex-
Received in revised form 29 January 2014 ternal agitation in the presence of a gravitational field. Thus, operation of many of these devices is prone to
Accepted 31 January 2014
segregation. Many studies of granular flow have focused on gravity driven chute flows owing to their practical
Available online 10 February 2014
importance in granular processing and to the fact that the relative simplicity of this type of flow allows for devel-
Keywords:
opment and testing of new theories/equipment. In the present work, we observe the deposition behavior of both
Granular materials mono-sized and polydisperse dry granular materials in an inclined chute flow, both experimentally and compu-
Chute flow tationally. We investigate the effects on the mass fraction distribution of granular materials of different parame-
Mixing ters such as chute angle, particle size, falling height and charge amount. The simulation results obtained using the
Segregation Discrete Element Method (DEM) are compared with the experimental findings and a high degree of agreement is
DEM observed. Tuning of the underlying contact force parameters allows realistic results and is used as a means of
validating the simulation model against available experimental data. The tuned simulation is then used to test
predictions of granular segregation theories outlined in previous studies. That is, it has been predicted that a
system can maintain a mixed configuration when L b Uavgts, where L, Uavg, and ts denote the length of the
chute, the average stream-wise flow velocity of the particles, and the characteristic time of segregation, respec-
tively. Our results with the tuned simulation support these conclusions for chute flows.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction there exists an inhomogeneous distribution of the charge – which con-


sists of materials of various sizes with different physical properties – as
Granular materials are ubiquitous and it is hard to find a process this leads to inconsistencies in the pressure drop across the cross section
industry which does not handle granular materials in some form. of the blast furnace. Since the hot gas flowing through this packed bed of
These materials are widely used in chemical, pharmaceutical, cement, materials actually reduces the metallic charge (iron ore) and produces
food, construction and metallurgical industries. When a mixture of hot metal or liquid iron (which is later converted to steel), segregation
two different types of particles is processed in any solids handling plays a major role in determining the efficacy of the reaction due to
device, e.g., in a chute, they tend to segregate or de-mix if there is any the fact that both maldistribution of reactant materials and/or gas
difference in mechanical properties. In fact, even a “homogeneous” flow will tend to be detrimental. In many other industries, a similar
material will segregate if it does not possess a very narrow distribution case for the importance of the operation of chutes can be made. Never-
of mechanical properties (such as particle size). These undesired theless, there is often little to no connection between experimental ob-
phenomena cause problems in maintaining uniform product quality servations and segregation theories even for this simple device.
and can result in devastating revenue losses in some industries. In the present work, we attempt to show a direct connection with
As an example of the ubiquitous importance of chute-related segre- recently developed segregation theory and the operation of granular
gation, here we briefly describe the importance of understanding of seg- chute conveyors. As such, we examine the deposition behavior of both
regation in chute flows in a steel industry. Traditionally, metallurgical mono-sized and polydisperse dry granular materials in an inclined
industries have used chutes to transfer and distribute granular materials chute flow (quasi 2D) using both experiments and simulation (using
(such as coke, iron ore, etc.). The distribution of coke, ore, sinter and the Discrete Element Method – DEM). We perform a systematic study
other raw materials at the stock level of a blast furnace is very crucial to observe the effects of different operating parameters such as chute
for its smooth operation. In an attempt to insure a uniform distribution angle, particle size, falling height and charge amount on the mass frac-
of materials, a rotating chute (see Fig. 1) is used, whereby the rotation tion distribution of granular materials after deposition. Tuning of the
rate (rpm) and inclination angle can be adjusted in order to facilitate ho- underlying contact force parameters of the DEM allows us to achieve re-
mogeneous flow. A significant problem for steel processing arises when alistic results in the simulation model, thus we are then able to use the
tuned simulation model to find the critical chute length for segregation
⁎ Corresponding author. Tel.: +1 412 624 7362; fax: +1 412 624 9639. [22,23] for a few simple combinations of material properties. It is impor-
E-mail address: jjmcc@pitt.edu (J.J. McCarthy). tant to note that the lab-scale experimental apparatus is only 600 mm in

0032-5910/$ – see front matter © 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.powtec.2014.01.092
T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139 127

showed that in a chute flow larger particles rise to the top of


the layer which is also in agreement with these results [4].
Dolgunin et al. [7,11] studied segregation in a chute flow for
both size and density. For size segregation they used a tight
size range (6.6–7.0 mm smooth steel balls), while the density
segregation experiments were performed for nearly a 1:2 density
ratio of two different materials. The interesting result from their
work is that they obtained a nonmonotonic (S shaped) concentration
profiles for different components. They also validated their continu-
um mathematical model of segregation by obtaining very good
agreement with experiments. Though a model involving the consti-
tutive equation for segregation flux was developed earlier by
Khakhar et al. [21] for density segregation in a tumbler using an ef-
fective medium approach, the work of Dolgunin et al. appears to be
one of the first continuum models of segregation involving a chute
flow, for both size and density. Later, Khakhar et al. [13] noted that
the model of Dolgunin et al., however, is not rigorous and thus
does not clearly specify the driving forces for the segregation or the
dependence of the segregation flux on the particle properties. The
work of Khakhar et al. [13] and Ottino et al. [14] critically reviewed
the different segregation models and observed that, although the
theories for segregation provide some physical insight into the pro-
cess and are reasonably successful in describing segregation in
chute flows, few are grounded on fundamentals. They suggested
that statistical mechanical studies (based on kinetic theory) of
hardsphere mixtures can provide a starting point for understanding
granular segregation. They noted that the kinetic-theory results permit
a general understanding of the causes of segregation. They were suc-
cessful in developing the segregation fluxes based on number fraction
or concentration gradient (ordinary diffusion), granular temperature
gradient (temperature diffusion) and granular pressure gradient (pres-
sure diffusion). While the present work does not set out to eluciate
proper modeling techniques for segregation rates, it does rely on cur-
rent (kinetic-based) theories to inform the analysis of our results. As
such, it is useful to define the idea of granular temperature, as it is
used here, since this concept is central to some of the models explored
in this work. Granular temperature is proportional to the kinetic energy
of the velocity fluctuations of the particles and is defined as:
Fig. 1. Example of chute flow in actual practice. An industrial chute is used for transfer and
distribution of granular materials on top of a blast furnace for producing liquid iron in a  
2 2
steel plant. This picture shows a bell-less top (BLT) charging device. T ¼ bu N−buN ð1Þ

where u is the velocity and the angled bracket indicates a time average.
length; therefore, the tests of critical segregation length and/or time are
3. Segregation modeling and mitigation
accomplished using the tuned/validated simulation.

Time-modulation in fluid mixing and other dynamical systems


2. Background [24] is a fairly common practice, but has found only limited ap-
plications in granular processing [25–27]. Recently, Shi et al.
A literature search reveals that there have been many studies [22] have suggested that segregation can be eliminated/mitigated
[1–20] of granular processing which focus on chute flows both if a flow is perturbed above a critical forcing frequency via peri-
due to the practical importance of such flows in granular trans- odic flow inversion. An alternative to this method, which is ex-
port as well as the relative simplicity of this type of flow as a plored in detail here, is to instead halt the flow at a time
model for detailed testing of theory [13]. Savage and Lun [4] ob- shorter than the critical time (period) of segregation. The key
served that in a chute flow, for high solid volume fractions, large to adapting this idea to free-surface segregation lies in recogniz-
voids are less likely to be formed than small voids. As a result, ing two important facts: that it takes a finite time for material to
smaller particles percolate through the voids within the larger segregate and that there is always a preferred direction that par-
particles to layers below, resulting in a net segregation flux of ticles tend to segregate. For example, in a free-surface flow (as
the smaller particles in the downward direction (i.e., normal to discussed above), small particles need time to percolate through
the chute surface or inclined plane). The above authors [4] the flowing layer; thus, if the flow is interrupted before the small
presented experimental data for polystyrene beads of varying particles could reach the bottom of the flowing layer, segregation
sizes in a chute flow, compared the distance along the chute at can be prevented. This relatively simple observation can be
which the material segregated completely into two layers and employed to engineer systems that counteract segregation, and
found their values to be in agreement with theory. Hirschfield can be tested directly in a chute flow segregation tester using
and Rapaport [10], using Molecular Dynamics simulations, scaling arguments as explained below.
128 T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139

The central point of the model is that one can estimate the value of Using these expressions, we can write a form for the characteristic
the characteristic segregation time tS and the critical forcing frequency segregation time as
fcrit which would ultimately aid in developing the expression for critical h  i
chute length. Recently, Hajra et al. [23] have proposed the following t s ¼ R1 = ðK T þ ð1−cÞK S Þð1−cÞ 1−d ; ð6Þ
form of the segregation velocity for a size segregating species:

  where R1 is the radius of the small particles. Now using this value, we
vs ¼ −Kd2 ð1−cÞ d−1 ð2Þ can define a segregation-based Peclet number by defining a diffusion
time-scale as R21/D so that we get
where d ¼ d1 =d2 is the size ratio or particles of types 1 and 2 (whose  
diameters are given as di), and c is the mass fraction of the smaller ðK T þ ð1−cÞK S Þð1−cÞ 1−d R1
Pe ¼ ; ð7Þ
particles. Assuming that the constant K has both an intrinsic and a D
concentration-dependent component (KT and KS, respectively) that
can be considered complex functions of granular temperature, local where D is the collisional diffusivity. Because of the current theoretical
void fraction, gravity, particle sizes, density, shape, roughness, coeffi- uncertainty and the time-varying nature of our flow (as well as our
cient of friction, coefficient of restitution, etc. we get granular temperature, local void fraction, system non-uniformity, etc.),
we treat β = KTR1/D and α = KSR1/D as fitting parameters that should
  be a decreasing function of fluctuation energy of the flow and should
vs ¼ ½K T þ ð1−cÞK S ð1−cÞ 1−d : ð3Þ be close to unity at small to moderate energies. This yields
 
Here the parameters, KT and KS, are considered to be fitting constants Pe ¼ ½β þ ð1−cÞα ð1−cÞ 1−d : ð8Þ
that are obtained on a case-by-case experimental/computational basis.
Using Eq. (3), we can obtain a segregation flux of the segregating species Finally, the particle diffusivity in sheared granular flows was obtain-
as ed by Savage [28] from numerical simulations of shear flow of nearly
2
elastic hard spheres to yield a scaling of the form D ¼ F ðν Þd γ̇, where d
J s ¼ vs c: ð4Þ is the particle diameter,γ̇is the shear rate, and F(ν) is a function of the
solid volume fraction (ν) (Hajra and Khakhar [29] confirmed the scaling
By substituting for vs from Eq. (3), we get experimentally).
By using the diffusivity as given by Savage [30] (D ¼ 0:01R21 γ̇), we
 
get ts written as
J s ¼ ½K T þ ð1−cÞK S  1−d cð1−cÞ: ð5Þ
2
tD R 100
ts ¼ ¼ 1 ¼   ; ð9Þ
Pe DPe ½β þ ð1−cÞα ð1−cÞ 1−d γ̇

where γ̇ is the shear rate. Finally, taking the flow down an inclined
chute to be essentially linear, we can write γ̇¼ 2U avg =H, where Uavg is
the average stream-wise flow velocity and H is the height of particle
stream flowing down the chute at a stable region. Therefore, by using
the expression U avg ¼ γ̇H=2, the critical chute length to initiate segrega-
tion is then given by

100U avg
Lcrit ¼ U avg t S ¼  
½β þ ð1−cÞα ð1−cÞ 1−d γ̇
50H
¼h : ð10Þ
β þ ð1−cÞαð1−cÞð1−d

While any number of modified models of segregation velocity (and


particle diffusivity) may be used, the critical issue of this section is
that we can relate measurable flow and particle parameters to a predict-
ed length over which particles will need to flow prior to macroscopic
segregation becoming noticeable. We should further note that the
types of mixtures that we can examine in this type of work, at present,
are limited by the availability of segregation velocity models. Hence, our
work in this paper is limited to a relatively simplistic mixture of three
differing particle sizes.

4. Methodology
Fig 2. Schematic of the chute flow experimental setup showing a model hopper, chute and
deposit bins (segregation box). The chute length is 600 mm, the vertical distance between As mentioned in the Introduction section, in this work we use both
the chute hinge and the hopper opening (flow control gate) is 100 mm, the hopper cross experimental measurements of chute flow segregation (or lack thereof)
section is square (150 mm × 150 mm); the setup is quasi-2D with a depth of 150 mm and and compare these quantitatively with simulations of the same device
a width of 1200 mm (1:8 aspect ratio). There are 20 bins each having a volume of 60 mm ×
60 mm × 150 mm (L × H × W). The bins can take up to three different vertical positions
(and materials). After affirming this validation, we perform further sim-
with respect to the chute hinge. Levels 1, 2 and 3 as described in the text are 800 mm, ulations to establish a wider range of parameter space for this study
1050 mm and 1300 mm below the chute hinge, respectively. (both in terms of material properties and effective device geometry).
T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139 129

4.1. Experiment

The experimental results are obtained on a test unit [31] as


shown in Fig. 2. The setup consists of an inclined chute of rectangu-
lar cross section made up of PMMA (polymethylmethacrylate) in a
quasi-two-dimensional rectangular cell. The depth to width ratio is
1:8. The bottom of the chute is rough and it can be fixed at two dif-
ferent angles (45° and 60°) to the vertical. To control the depth of
the moving bed along the chute, a pneumatic control gate is
installed at the hopper opening to regulate the flow. The gate open-
ing pressure is varied between 2 kg/cm 2 –5 kg/cm 2 , which causes
the gate to open essentially instantly. A horizontal tray (called a
“segregation box”) with 20 bins is used to collect the deposited ma-
terials and can be installed in a rail inside the setup at three differ-
ent levels (heights). Nearly spherical polystyrene particles of
different sizes (6, 7 and 14 mm) are used as the test granular mate-
rial. A vacuum cleaner is used to collect the particles from the dif-
ferent bins of the segregation box. The collected particles are later
weighed using a highly accurate electronic weighing balance (Sar-
torius make, Model GP 3202, capacity 3200 g with 0.01 g accuracy).

4.2. Modeling: Discrete Element Method (DEM)

The Discrete Element Method (DEM), also known as Particle Dy-


namics (PD), is a simulation technique whereby the trajectory of each
particle within a system is tracked via simultaneous integration of the
interaction forces between individual pairs of particles [32,33]. While
these forces typically include only contact forces – normal (Hertzian)
repulsion and tangential (Mindlin) friction, see Ref. [34]. – and gravity, Fig. 3. The chute flow setup as used in the numerical simulations. This typical snapshot
additional particle interaction forces (such as surface adhesion [35,36], shows 14 mm diameter particles flowing down a 600 mm chute inclined at 60° with the
vertical. Hopper with the gate (opened to the right, short black line) is also shown in
van der Waals [37], and even heat transfer effects [38,39]) can be easily
this picture. The chute, bins and other fixed walls are made from 4 mm diameter particles.
incorporated. In particular, this technique has been quite successful in There are 20 bins spanning the full width of the setup (1200 mm). The simulation is peri-
simulating ensembles of granular materials [40], yielding insight into odic in the direction pointing into the plane of the paper. Bins are numbered from 1 to 20
such diverse microscopic phenomena as force transmission [41], pack- (R to L): bins 1–10 are called wall-side bins and bins 11–20 are called chute-side bins.
ing [42], wave propagation [43], agglomeration formation and breakage
[35], cohesive mixing [44,45], bubble formation in fluidized beds [46], such as chute angle, particle size, falling height and charge
and segregation of free-flowing materials [47,21]. In this work, the col- amount on the mass fraction distribution of granular materials
lisional forces are modeled after the work of Hertz and Mindlin [34]. A after deposition. Starting with the mono-disperse case, we first
thorough description of the interaction laws from contact mechanics tune the underlying contact force parameters of the DEM model
(collision forces) can be found in Refs. [48–52]; therefore, they are not so that we can obtain agreement between the experimental and
reviewed here. computational results. A sensitivity analysis to what is deemed
Typically, a numerical experiment consists of a mono-disperse/ the most significant model parameter, namely the yield stress
polydisperse system of perfectly smooth spheres bounded by a wall of im- of the material, is performed so that appropriate contact force
mobile particles with periodic boundaries in the z (depth) direction. The parameters can be chosen to validate the simulation.
simulation parameters (chute angle, particle size, etc.) for various cases In the second part of this section, the tuned simulation is used
considered in the present study were inspired by many experiments to find the critical chute length for segregation. This is done both
[31] performed on an identical setup (refer to Fig. 2 in Section 4.1). for a batch/finite length chute as well as for a periodic (infinite
Fig. 3 shows the simulation setup which closely resembles the actual length) chute. The values obtained are compared with the
experimental setup. The simulated particle sizes and densities as well as length/time that is predicted based on a recently proposed theo-
apparatus dimensions were matched as closely as possible to their corre- ry by our group (as discussed in Section 3).
sponding experiments so that results can be directly compared. A typical
simulation evolves with particles starting from rest in the hopper after 5.1. Mono-disperse flow: simulation vs. experiment
gravitational settling. The hopper gate is opened and as time elapses,
the particles move under the influence of gravity resulting in their flow 5.1.1. Tuning of the contact force model
through the chute and finally their deposition in various bins after leaving As mentioned, we first use the mono-disperse experimental results
the chute (refer to Fig. 3). as a means of tuning our modeling efforts. We considered two force
models in our PD simulations: a spring-dashpot (elastic with viscous
5. Results and discussion
Table 1
Material properties used in the simulations (and experiments).
In this section, we discuss the results relating to both the deposition
behavior of both mono-sized and polydisperse dry granular materials in Parameter Value
a chute flow as well as those comparing our simulations to predictions Young's modulus (E, GPa) 2.9
of the critical length/time for segregation. Specifically, in the first part Density (ρ, kg/m3) 951–1160
of this section we compare the results obtained from experiments and Coefficient of friction (μ) 0.30
DEM simulations for the flow of a stream of granular material down Poisson ratio (ν) 0.33
Yield stress (σy, MPa) 45.0
an inclined chute. We also examine the effects of different parameters
130 T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139

a b

Fig. 4. Comparison of simulation results from two force models with the data obtained from the experiments. The force models PD Elastic and PD Plastic denote elastic with viscous damping
and elasto-plastic models, respectively. (a) corresponds to a case with 7 mm diameter polystyrene balls (1 kg) with a 60° chute and the bins placed at level 1. (b) corresponds to 6 mm
diameter polystyrene balls (0.537 kg) with a 45° chute and the bins placed at level 1. Refer to Fig. 2 for the positions of different levels.

dissipation) model and an elasto-plastic model (see [40] for details). The peak position and peak height for the PD Plastic model with a
They are also referred to as PD Elastic and PD Plastic, respectively, in yield stress of 45 MPa are bin number 7 and 27.96%, respectively.
this paper. Table 1 shows the material properties that are used in the This is close to the experimental observation of peak position at
simulations. Fig. 4(a) and (b) compares the results from the spring- bin number 6 and peak height at 28.65%. The corresponding values
dashpot and elasto-plastic force models with the experimental data for the next closest case (PD Plastic model with 22.5 MPa yield
for two different experimental cases as mentioned in the caption. stress) are bin number 7 and 31.69%, respectively. Note that here
Though the simulation results agree well with the experiments for we prefer peak height over peak position to select the model as
both types of force models, we chose the elasto-plastic force model be- both the yield stresses give a close peak position when compared
cause of its superior capability to capture the underlying physics and with the experiment. In addition to the above comparisons, we
due to the fact that all the model parameters can be related to also perform a two-sample Kolmogorov–Smirnov (K–S) test [53]
measureable material properties. In contrast, the spring-dashpot model to cross-check the agreement between the experimental result
uses only the empirical damping parameter that cannot be obtained and PD Plastic models with different yield stress parameters. The
from material properties. two-sample K–S test is generally performed for comparing two
Next, we tune the yield stress used in the PD Plastic force samples or distributions, as it is sensitive to differences in both lo-
model to mimic realistic contact mechanics. The tuning is per- cation and shape of the empirical cumulative distribution functions
formed in the following way: we adjust the yield stress in order (CDFs) of the two samples. The two-sample K–S test returns the
to match a single set of experimental data and then the adjusted probability (p) of observing the given statistic (i.e., whether the
yield stress is kept fixed in the model and is subsequently used to two distributions are from the same continuous distribution or
compare results from other experimental data or make predic- not – a p value of 1.0 signifies that the two distributions are identi-
tions for different chute flow cases. Fig. 5 shows that a yield cal) and also quantifies a distance (k) between the empirical distri-
stress of 45 MPa reproduces the experimental results very well. bution functions of the two distributions. A lower k value signifies a
better agreement between the two distributions. Table 2 summa-
rizes the results of the two-sample K–S test performed on the
yield stress tuning data. The experimental data are used as one of
the samples for all cases. One can clearly observe that the experi-
mental data is best matched by the PD Plastic force model with a
yield stress of 45 MPa (which we refer to as 1x in our sensitivity
analysis) as it produces the highest p value and the lowest k
value. Therefore, we select the PD Plastic model with 45 MPa yield
stress to simulate various cases as described in the next sub-
sections (coincidentally, the yield stress of polystyrene beads is
∼ 45 MPa, which indicates that our PD Plastic model has a sound
force model and captures the contact mechanics very well).

Table 2
Two-sample Kolmogorov–Smirnov (K–S) test results of elasto-plastic force model tuning
data.

Case p value k value

PD Plastic (YS: 1/2x) vs. Experiment 0.50 0.25


PD Plastic (YS: 1x) vs. Experiment 0.97 0.15
Fig. 5. Tuning of the elasto-plastic force model parameter (YS: yield stress) for a case sim-
PD Plastic (YS: 2x) vs. Experiment 0.50 0.25
ilar to the experimental condition as in Fig. 4(a). A base YS value of 1x corresponds to
PD Plastic (YS: 10x) vs. Experiment 0.00 (test failed) 0.50
45 MPa.
T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139 131

Table 3
Different cases for mono-disperse simulation (and experiment) and the operating param-
eters used.

Case Operating parameters

Case 1: Effect of Mass of particles: 0.3 kg (N = 1455 particles),


charge amount 0.6 kg (N = 2910), 0.9 kg (N = 4365)
Diameter of particles: 7 mm
Chute angle: 45°
Bin location: level 2 (1050 mm below chute hinge)
Case 2: Effect of Diameter of particles: 7 mm and 14 mm
particle size Mass of particles: 1 kg (N7 = 4850, N14 = 600)
Chute angle: 60°
Bin location: level 1 (800 mm below chute hinge)
Case 3: Effect of Bin location: levels 1, 2 and 3
falling height (level 3 is 1300 mm below chute hinge)
Diameter of particles: 6 mm
Mass of particles: 1 kg (N = 5000)
Chute angle: 60°
Case 4: Effect of Chute angle: 45° and 60°
chute angle Diameter of particles: 6 mm
Mass of particles: 0.537 kg (N = 5000)
Bin location: level 1

Fig. 7. Effect of particle size: comparison with experimental data.

obtained, confirming that there is a good agreement between experi-


In the following sub-sections, simulation results for the various cases
ment and simulation. The main observation from this graph is that
mentioned above are presented and whenever applicable, comparison
there is no significant effect of the initial charge amount on where
is made with the experimental data. The various cases of simulations
most of the particles get deposited after leaving the chute; however,
and the operating parameters are listed in Table 3.
there is some effect on the peak height or the maximum mass fraction
of particles.
5.1.2. Case 1: effect of charge amount (mono-disperse)
The only parameter which is varied for this case is the total number
5.1.3. Case 2: effect of particle size (mono-disperse)
(or mass) of particles in the hopper. This is also known as the charge
Two different sized particles, 7 mm and 14 mm, are considered sep-
amount in industrial practice. Refer to Table 3 for a complete description
arately. Fig. 7 shows the comparison between modeling results and the
of all the operating parameters. The hopper is filled with three different
experimental data. We observe that the similarity between these two is
amounts of particles separately and in each run, the particles are
significant (the two-sample K–S test gives a very high p value of 96.5%
allowed to fall on the chute. Fig. 6 shows the comparison between sim-
for both particle sizes), yet there is a slight mismatch at the wall
ulation and experiment of the effect of charge amount on mass fraction
side (bins 1–3) for 14 mm particles. However, the overall trend is
distribution.
comparable.
All peak positions are centered on bin number 7 and the agreement
If we closely observe the distributions, we notice that despite dou-
becomes better for higher charge amount. The peak heights are 26.72%
bling the particle size, the effect on the final distribution is not very sig-
(300 g, simulation), 35.02% (300 g, experiment), 31.10% (600 g, simula-
nificant. Also, the location of the densest region in the trajectory is
tion), 38.31% (600 g, experiment), 32.28% (900 g, simulation) and
nearly insensitive to the size of the particles.
33.37% (900 g, experiment). Clearly, there is a qualitative agreement be-
tween the trends for both simulation and experiment. In addition to
comparing peak positions and peak heights of the distributions we, 5.1.4. Case 3: effect of falling height (mono-disperse)
again, perform a two-sample K–S test and confirm that the two distribu- In this case, the falling height of the particles is varied at three differ-
tions (experiment vs. simulation) are from the same continuous distri- ent levels (see Table 3). Fig. 8 shows three snapshots of the particle tra-
bution. A high probability (p = 96.5%) of observing the given statistic is jectory corresponding to three different falling heights. Fig. 9 offers the

a b c

Fig. 6. Effect of charge amount: comparison with experimental data. (a), (b) & (c) correspond to 300, 600 & 900 g of materials respectively.
132 T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139

Fig. 8. Snapshots showing trajectories for three different falling heights.

comparison between experiment and simulation of the mass fraction angles, respectively. As expected from Fig. 10, a higher chute angle pro-
distribution for each of the falling heights. We can observe that, as ex- duces a wider trajectory with the peak (densest region) shifted towards
pected, the densest region of the trajectory (highest mass fraction or the wall.
peak of the distribution) shifts towards the wall side as the falling height
is increased and the densest region also gets thinner (peak height de-
creases). The simulation accurately captures the right bin number for 5.2. Polydisperse flow: simulation vs. experiment
the maximum deposition for level 1 and level 2. There is some mismatch
when the falling height is increased to level 3. The maximum % of mass In this section, we discuss the experimental and simulation re-
fraction, for the case of level 1 in experiment is 30.84% (24.66% for level sults for chute flow for a mixture of particles with different sizes.
2 and 17.67% for level 3) and the corresponding simulation value is First, two cases are analyzed: The effect of falling height and of
30.1% (23.92% for level 2, and 19.78% for level 3). The two-sample K–S chute angle. By validating that our experiments and simulations
test gives a (p, k) value of (0.99, 0.10), (0.77, 0.20) and (0.77, 0.22) for agree for the results of polydisperse mixtures, we set the stage for
levels 1, 2 and 3, respectively, when compared with experiments. the direct measurement of the critical chute length/time (via sim-
Therefore, we can conclude that the simulation captures the essential ulation) in the next section.
features (peak position, peak height and distribution shape) of the In all cases examined here, the size difference between the differing
chute flow very well for the lowest falling height and to a reasonable ex- combinations of particles is less than the spontaneous percolation
tent for increased falling heights. threshold [54,55]. Therefore, we expect that the observed segregation
is dominated by a shear-induced percolation mechanism. The various
cases and the operating parameters of both the simulations and the ex-
5.1.5. Case 4: effect of chute angle (mono-disperse) periments are listed in Table 4. The same tuned simulation from the
For this case, the chute angle is varied at two different levels: 45° and mono-disperse case is used for all of the reported simulations.
60°, and all other parameters are kept constant. There is a high degree of To match the experimental conditions, the particles are randomly
agreement between simulation and experiment as is evident from the placed in the hopper to generate a homogeneous mixture as this is
visual observation. We also perform the two-sample K–S test and obtain what was used in the experimental hopper trials. To ascertain that the
a (p, k) value of (0.97, 0.15) and (0.99, 0.10) for 45° and 60° chute initial randomness does not affect the final mass distribution in the
bins, some sensitivity tests were run before actual simulations were
50
level 1 (Model)
50 45° (Model)
level 1 (Expt)
40 45° (Expt)
level 2 (Model)
60° (Model)
Mass Fraction (%)

level 2 (Expt) 40
60° (Expt)
Mass Fraction (%)

30
level 3 (Model)
level 3 (Expt) 30

20
20

10
10

0
2 4 6 8 10 12 14 16 18 20 0
Bin No. 2 4 6 8 10 12 14 16 18 20
Bin No.
Fig. 9. Comparison of simulation and experimental results for observing the effect of falling
height: level 1 is 800 mm below chute hinge, level 2 is 1050 mm below chute hinge and Fig. 10. Effect of chute angle: the chute is fixed at two different angles, 45° and 60°, by
level 3 is 1300 mm below chute hinge. keeping other parameters unchanged.
T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139 133

Table 4 respectively, and the bins are kept at a fixed elevation of level 2 in
Different cases for polydisperse simulation and operating parameters. both the cases. This controlled experiment singles out the effect of the
Case Operating parameters chute angle on mass fraction distribution and particle size distribution
in a multi-sized mixture. Again, the individual distributions correspond-
Case 1: Effect of falling height Bin location: levels 2 and 3
Diameter of particles: 6 mm, 7 mm and 14 mm ing to simulation and experiment agree quite well. Additionally, we ob-
Mass of particles: 6 mm: 0.5 kg (N = 4641), serve that, like the mono-disperse case, the qualitative shifts in the
7 mm: 0.3 kg (N = 1455), 14 mm: 0.2 kg (N = 119) peaks agree well. That is, as the angle increases, all peaks shift towards
Chute angle: 45°
the wall side (lower bin numbers) and the corresponding height of
Case 2: Effect of chute angle Chute angle: 45° and 60°
Diameter of particles: same as Case 1 the peaks also decrease.
Mass of particles: same as Case 1
Bin location: level 2 5.3. Critical segregation length and time

performed. To do this, the initial arrangement of particles was generated Following the discussion of the critical chute length found in
using various random number generators (e.g., gsl_rng () from GNU Sci- Section 3, we now attempt to quantify this value for the chute setup
entific Library [56], drand48(), rand() and random() from standard C li- under consideration using our tuned/validated simulations. In other
brary) and it was found that the final mass fraction distribution did not words, by varying the length of the chute, we can effectively measure
depend on the random number generator used (i.e., on the initial the critical chute length for segregation of the mixture of particles con-
randomness). sidered in this study.

5.3.1. Critical chute length for segregation


5.2.1. Case 1: effect of falling height (segregating)
Fig. 13(a) and (b) shows how the velocity profile of particles on a
Fig. 11 shows the mass fraction distribution of differently-sized par-
stable region of chute is calculated. This information is needed to obtain
ticles after they are mixed randomly in the hopper and allowed to flow
the shear rate (γ̇) in the system.
over the chute. The chute angle is fixed at 45° and the elevation of the
Starting with a base case as shown in Fig. 13(a), we systematically
bins is varied at level 2 and level 3, respectively. This controlled experi-
vary the simulated chute length from 300 mm to 7000 mm by keeping
ment singles out the effect of the falling height on mass fraction distri-
other parameters (such as falling height, chute angle, etc.) unchanged.
bution and particle size distribution in a multi-sized mixture. The
Then we monitor the mass fraction distribution for each of the simula-
individual distributions corresponding to the simulation and experi-
tion cases. A narrow distribution for all the particle sizes would mean
ment show a significant match for most species under most conditions.
that all the particles with different sizes have been deposited within a
The only exception is that there is some mismatch for the largest parti-
small region spanning only a few bins, and we consider this to be a
cle size at the highest falling height. Nevertheless, even in this case both
case similar to mixing of particles (think of a mixture of particles in a
simulation and experiment indicate a similar trend—the peaks shift to
container or bin).
the wall side (towards lower bin numbers) as the falling height in-
A wider distribution for one particle size, on the other hand, would
creases. Taking all of the results in composite, it is evident that in all
mean that there is considerable spread of that type of particles and
cases the larger particles tend to deposit towards the wall side (i.e.,
the deposition would span across multiple bins. We consider this to
lower bin numbers).
be akin to segregation as all types of particles are no longer deposited
in one small region. Essentially, a narrow distribution for one particle
5.2.2. Case 2: effect of chute angle (segregating) size and a wider distribution for another particle size is a typical signa-
Fig. 12 shows the mass fraction distribution of multi-sized particles ture for segregation. Therefore, segregation is deemed to occur if various
for two chute angles. The chute angles are varied at 45 and 60°, particle sizes have differing degrees of spread in the mass fraction

60 Expt (6 mm, level 2)


60
Expt (7 mm, level 2) Expt (14 mm, level 2)
Model (6 mm, level 2) Model (7 mm, level 2) Model (14 mm, level 2)
50 50

40 40

30 30

20 20
Mass Fraction (%)

10 10

0 0

60 Expt (6 mm, level 3) Expt (7 mm, level 3) Expt (14 mm, level 3)
Model (6 mm, level 3) Model (7 mm, level 3) Model (14 mm, level 3)
50
40
30
20
10
0
2 4 6 8 10 12 14 16 18 20 2 4 6 8 10 12 14 16 18 20 2 4 6 8 10 12 14 16 18 20
Bin No.

Fig. 11. Effect of falling height on mass fraction distribution of a mixture of polydisperse particles. Both experiments and simulation results have been shown for each particle size and
falling height as mentioned in the legend.
134 T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139

60 Expt (6 mm, 45o)


60
Expt (7 mm, 45o) Expt (14 mm, 45o)
Model (6 mm, 45o) Model (7 mm, 45o) Model (14 mm, 45o)
50 50

40 40

30 30

20 20
Mass Fraction (%)

10 10

0 0

60 Expt (6 mm, 60o) Expt (7 mm, 60o) Expt (14 mm, 60o)
Model (6 mm, 60o) Model (7 mm, 60o) Model (14 mm, 60o)
50
40
30
20
10
0
2 4 6 8 10 12 14 16 18 20 2 4 6 8 10 12 14 16 18 20 2 4 6 8 10 12 14 16 18 20
Bin No.

Fig. 12. Effect of chute angle on mass fraction distribution of a mixture of polydisperse particles. Both experiments and simulation results have been shown for each particle size and chute
angle as mentioned in the legend.

distribution. This is evident from Fig. 14 where we can observe that as the 6 mm (or 7 mm) particles are distributed over a wider distance
the chute length increases, the distribution becomes wider (i.e., peak (from about 8 bins or 480 mm for chute length 300 mm to about 25
height reduces as the sum of the distributions is 100%) for smaller par- bins or 1.5 m for chute length 7 m). This signifies that as the chute
ticles whereas larger particles continue to have a sharper peak. In this length increases, there is more and more separation occurring between
figure, to have a comparison on a uniform distance scale for all particle particles with different sizes. Therefore, from this distribution plot, we
sizes and chute lengths, all the peak positions corresponding to 14 mm can get some idea about how segregation can be controlled by varying
particle distribution have been rescaled so that it is denoted as bin num- the chute length.
ber 0 (an arbitrary choice). The negative bin numbers correspond to the Now, turning our focus back to the establishing of the critical chute
bins to the left of the 14 mm peak position (towards the wall side) and length to initiate segregation, we tabulate the pertinent parameters of
the positive bin numbers correspond to the bins to the right (i.e., to- Eq. (10) for our case in Table 5.γ˙ has been calculated from the slope of
wards the chute side). The important observation from these plots is the linear portion of the velocity profile in Fig. 13(b). The equation
that 6 mm and 7 mm particle sizes have a wider mass fraction distribu- gives a value of 4.0 m as the critical chute length for the present system
tion as compared to 14 mm particles for higher chute lengths. We ob- under consideration. Now, if we plot the peak height (i.e., maximum
serve that the 14 mm particles are always distributed over a relatively mass fraction) for all the distributions (for all particle sizes) as a function
narrow distance (about 4 bins or 240 mm) for all chute lengths, whereas of chute lengths, we obtain a plot as shown in Fig. 15(a). It is evident

a b

Fig. 13. (a) A typical flowing layer on a 600 mm chute with 45° inclination. The bins are placed at level 2 (1050 mm below chute hinge). The bed of particles consists of 6, 7 and 14 mm
particles in a weight ratio of 5:3:2, respectively. (b) Velocity profile of particles is calculated on a small slice of width 28 mm centered at half chute length corresponding to the scenario
shown in (a). Situation (a) is chosen because of a fully developed layer with uniform thickness around the middle of the chute.
T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139 135

60 14 mm (L = 300 mm)
50 7 mm (L = 300 mm)
6 mm (L = 300 mm) L = 600 mm L = 900 mm L = 1800 mm
40

30

20

10

0
60
Mass Fraction (%)

50
L = 2400 mm L = 3000 mm L = 3600 mm L = 4200 mm
40

30

20

10

0
60

50
L = 4800 mm L = 5400 mm L = 6000 mm L = 7000 mm
40

30

20

10

0
-25 -20 -15 -10 -5 0 5 -25 -20 -15 -10 -5 0 5 -25 -20 -15 -10 -5 0 5 -25 -20 -15 -10 -5 0 5
Distance (bins)

Fig. 14. Mass fraction distribution of 14, 7 and 6 mm particles in a polydisperse chute flow for different chute lengths. To obtain a clear comparison, 14 mm particle peak positions for all
chute lengths have been arbitrarily set to bin number 0. Other conditions are similar to Fig. 13(a).

from this plot that as the chute length increases initially, the peak height completely mixed and R ≤ 0.25 (when the spread of small particles is
decreases rapidly and that after about L = 5.4 m of chute length, the three times the spread of larger particles) is taken (somewhat arbitrari-
peak height becomes nearly flat and does not change significantly. The ly) to signify segregation. This plot also confirms the fact that segrega-
peak height is a measure of the spread of the distribution, and therefore tion is initiated at about L = 5.4 m of chute length. Therefore, our
the degree of segregation. That is, a lower peak height corresponds to a theoretical prediction and the prediction from the computer model
wider spread, and a wider spread signifies segregation. Also, in are of the same order of magnitude (4.0 m vs. 5.4 m).
Fig. 15(b), we plot the degree of mixing (a new mixing measure specific
to chute flows), R (Eq. (11)), as a function of chute length.
Based on this argument regarding the peak spread/height, we define 5.3.2. Critical length for segregation: finite vs. periodic chute
the degree of mixing R as The theoretical arguments used to predict the critical chute length
assume a fully developed shear flow. As we have seen in the earlier sec-
R ¼ W 14 =ðW 6 þ W 14 Þ ð11Þ tions, this assumption may not remain valid if longer chutes are used
with a limited number of particles. In other words, a finite chute flow
where Wi is the full width at half maximum (FWHM) of the mass frac- is essentially a batch process where shearing of the flow will lead to
tion distribution for a particular particle size i (we consider 14 mm thinning of particle layers. Therefore, if a longer chute is used, this
and 6 mm, the largest and smallest particle sizes, respectively). A may cause the particles to segregate in the flowing direction and not
value of R = 0.5 (i.e., W14 = W6) indicates that the two sizes are along the normal-to-flow direction as assumed in the critical chute
flow argument. With this issue in mind, we realize that using a finite
Table 5 chute flow to identify the critical chute flow is strongly pushing the
Parameters to calculate critical chute length from limits of the simplistic segregation model set forth. As such, we also em-
Eq. (10). ploy a periodic chute flow setup to cross-examine whether our tests
Parameter Value from the batch flow obtain the same results as a more theoretically ap-
propriate flow (but one that is also obviously much more difficult to
β 2.0
c 0.7467 physically realize).
α 2.1 Before discussing the results of our periodic flow simulations, we
d 0.4286 first use our validated simulations to examine the degree to which our
Uavg 1.275 m/s finite chute flows do, in fact, exhibit flow-direction segregation. Fig. 16
γ̇ 22.0 s−1
shows how the average difference in centroid positions (along the flow
136 T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139

a b

Fig. 15. (a) Maximum mass fractions (peak heights) have been plotted as a function of chute length for a polydisperse chute flow simulation. Lower peak height corresponds to a wider
distribution and hence signifies considerable segregation. (b) Degree of mixing R has been plotted as a function of chute length.

direction) of bigger (14 mm) and smaller (6 mm) particles changes as a mm and the chute angle is set at 80° from the vertical. The choice for
function of time for chutes with different lengths. The y axis corre- the above angle is due to the fact that a chute angle of 45° (as used in
 
X c;big −X c;small the experiments and validation simulations) does not produce a steady
sponds to a quantity called X c , which is defined as X c ¼ ,
0:5Lplug flow in an infinite chute, as the particles continue to accelerate indefi-
where Xc is the centroid of the particle mass (for big or small) and nitely. Therefore, 80° is chosen as the chute angle for assessing segrega-
Lplug is the length of the particle plug on the chute at the time in ques- tion rates in the periodic chute case.
tion. This quantity measures the amount by which the two types of par- In order to calculate the theoretical characteristic segregation time
ticles are separated on the chute along the flow direction. The x axis tS, the shear rate (γ̇) is determined from the velocity profile at different
corresponds to the dimensionless time t as defined by t ¼ t t−t−t , exit
entry
entry
times during the flow and an average value is obtained from different
where tentry and texit correspond to the particle entry and exit time to/ time samples. Fig. 17 shows how the shear rate fluctuates over time
from the chute, respectively. A value of X c ¼ 1:0 signifies a complete and these data are used to find an average shear rate (γ̇). Table 6 tabu-
separation of the bigger and smaller particles on the chute along the lates all the pertinent data for calculation of tS (refer Eq. (10)). Note
flow direction. As is evident from this figure, there is some degree of that the fitting parameters α and β are kept the same as in the finite-
horizontal separation on chute for lengths above 4.8 m (the X c for length chute case and this yields a characteristic segregation time of
other intermediate chute lengths shorter than 4.8 m never exceeds about 7.2 s. Now, we investigate if we can arrive at this characteristic
0.5). Therefore, the assumption of solely vertical segregation on these segregation time – as obtained from the theory – from direct observa-
longer chutes is not completely valid justifying the need for further test- tion of the concentration profile that is observed in the validated simu-
ing using a periodic chute flow setup. lations. To accomplish this, first we plot the concentration profiles of all
In the periodic chute flow setup, all the simulation conditions are types of particles at different time instances in the periodic chute flow.
similar (particle numbers, etc.) to the finite chute length case except The bed of particles is sliced into many bins in the normal-to-flow direc-
that the periodic chute is enclosed in a simulation box with length 600 tion, and the volumes of all types of particles in a particular bin are used
to calculate the concentration of each kind of particle in that bin.

L = 600 mm
1.6 L = 900 mm 30
L = 1200 mm γ.
L = 1800 mm 25
1.2
L = 2400 mm
L = 4800 mm
L = 5400 mm 20
0.8
L = 6600 mm
L = 10000 mm 15
0.4

10
0

5
-0.4
0 0.2 0.4 0.6 0.8 1
0
0 10 20 30 40 50 60 70 80 90 100
Fig. 16. Normalized difference of the average centroid position of bigger and smaller par-
Time (s)
ticles on chute as a function of normalized time. Data for different chute lengths have been
plotted together. Fig. 17. The shear rate fluctuation over time in a periodic chute flow.
T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139 137

Table 6 Standard Deviation (RSD) (Fig. 20) and also (1 − ML) (Fig. 21), where
Parameters to calculate critical segregation time for a ML is the Lacey index. These indices are calculated considering 14 mm
periodic chute flow from Eq. (10).
particles as the tracer particles. Clearly, all these indices suggest
Parameter Value that after about 6.0 s (the vertical line in these graphs corresponds to
β 2.0 6.0 s), the degree of segregation in the periodic chute flow reaches an
c 0.7467 asymptotic value. Therefore, we can conclude that the direct observa-
α 2.1 tion of our simulations in a periodic chute flow gives a characteristic
d 0.4286
segregation time which is of the same order of magnitude as the predic-
γ̇ 9.61 s−1
tion from the theory (6.0 s vs. 7.2 s).

6. Conclusions & outlook


It is obvious from Fig. 18 that somewhere between 4 and 10 s, the
bigger particles (14 mm) start to separate. The concentration of 14 Granular materials are omnipresent. Many man-made or natural
mm particle becomes zero for about half of the bed height starting processes involve flow of granular materials. Industrial applications typ-
from time 4 s. To further narrow our identification of the time for segre- ically involve handling and processing of a large amount of multi-sized
gation, we plot three different indices for segregation as a function of granular materials which may have different shapes and densities.
time: the so-called Intensity of Segregation (IS) (Fig. 19), the Relative These processes require many solid handling devices like hoppers,

175
14 mm (0 s)
7 mm (0 s)
140 6 mm (0 s)
2s 4s

105

70

35

175

140 5s 7s 9s
Bed Height (bins)

105

70

35

175

140 10 s 12 s 14 s

105

70

35

0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Concentration (-)

Fig. 18. Evolution of concentration profiles in a periodic chute flow.


138 T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139

0.2 1.2

1
0.15
0.8
IS (-)

0.1 0.6

0.4
0.05
0.2

0 0
5 10 15 20 25 30 35 40 45 50 0 10 15 20 25 30 35 40 45 50
Time (s) Time (s)

Fig. 19. Evolution of Intensity of Segregation (IS) in a periodic chute flow. The vertical line Fig. 21. Evolution of the derived Lacey index (1 − ML) in a periodic chute flow.
demarcates the 6.0 s time after which the IS becomes flat.

chutes, etc. In this work, a simple flow of both mono-sized and polydis- Acknowledgments
perse spherical granular particles is analyzed when the particles were
allowed to fall from a hopper onto an inclined chute. The effect of vari- This work was supported by the following grants: ACS PRF
ous parameters like the charge amount, particle size(s), falling height 43877AC9, US DOE DEAC2604NT41817M2234 & NSF CBET0933358.
and chute angle were studied systematically to see how the mass frac- TB gratefully acknowledges Tata Steel for allowing him to publish the
tion distribution in the trajectories is affected by these parameters. A experimental results.
contact force parameter of the DEM model was tuned in order to obtain
reliable results and good agreement is obtained between simulations
and experimental observations. The tuned simulation is then used to References
find the critical chute length for segregation, and to test the efficacy of
[1] C. Campbell, Flow regimes in inclined open-channel flows of granular materials,
finite versus periodic chutes for segregation testing. Note that, while
Powder Technol. 41 (1) (January 1985) 77–82.
there was very little segregation evident in the lab-scale testing appara- [2] C.S. Campbell, C.E. Brennen, Chute flows of granular material: some computer sim-
tus (compare the results from Fig. 12 (top) and the first panel of Fig. 14), ulations, J. Appl. Mech. 52 (1985) 172–178.
extracting parameters that are experimentally accessible (such as the [3] M. Farrell, C. Lun, S. Savage, A simple kinetic theory for granular flow of binary mix-
tures of smooth, inelastic, spherical particles, Acta Mech. 63 (1) (November 1986)
shear rate) allows us to accurately predict how segregation will scale 45–60.
with increasing chute length. [4] S.B. Savage, C.K.K. Lun, Particle size segregation in inclined chute flow of cohesion-
less granular solids, J. Fluid Mech. 189 (1988) 311–335.
[5] Thorsten Poschel, Granular material flowing down an inclined chute: a molecular
dynamics simulation, J. Phys. II 3 (1) (January 1993) 27–40.
[6] O.R. Walton, Numerical simulation of inclined chute flows of monodisperse, inelas-
1 tic, frictional spheres, Mech. Mater. 16 (1993) 239–247.
[7] V. Dolgunin, Segregation modeling of particle rapid gravity flow, Powder Technol.
83 (2) (May 1995) 95–103.
[8] X. Zheng, Molecular dynamics modelling of granular chute flow: density and veloc-
ity profiles, Powder Technol. 86 (2) (1996) 219–227.
0.8 [9] Olivier Pouliquen, Nathalie Renaut, Onset of granular flows on an inclined rough
surface: dilatancy effects, J. Phys. II 6 (6) (June 1996) 923–935.
[10] D. Hirshfeld, D.C. Rapaport, Molecular dynamics studies of grain segregation in
sheared flow, Phys. Rev. E. 56 (2) (1997) 2012–2018.
0.6 [11] V. Dolgunin, A. Kudy, A. Ukolov, Development of the model of segregation of parti-
RSD (-)

cles undergoing granular flow down an inclined chute, Powder Technol. 96 (3)
(May 1998) 211–218.
[12] S. Dippel, D. Wolf, Molecular dynamics simulations of granular chute flow, Comput.
Phys. Commun. 121–122 (1999) 284–289.
0.4
[13] D.V. Khakhar, J.J. McCarthy, J.M. Ottino, Mixing and segregation of granular materials
in chute flows, Chaos 9 (1999) 594–610.
[14] J.M. Ottino, D.V. Khakhar, Mixing and segregation of granular materials, Annu. Rev.
Fluid Mech. 32 (2000) 55–91.
0.2 [15] L.E. Silbert, D. Erta\cs, G.S. Grest, T.C. Halsey, D. Levine, S.J. Plimpton, Granular flow
down an inclined plane: Bagnold scaling and rheology, Phys Rev E Stat Nonlin Soft
Matter Phys 64 (5 Pt 1) (November 2001).
[16] C. Ancey, Dry granular flows down an inclined channel: experimental investigations
on the frictional–collisional regime, Phys Rev E Stat Nonlin Soft Matter Phys 65 (1 Pt
0
5 10 15 20 25 30 35 40 45 50 1) (January 2002).
Time (s) [17] M.Y. Louge, Model for dense granular flows down bumpy inclines, Phys Rev E Stat
Nonlin Soft Matter Phys 67 (6 Pt 1) (June 2003).
[18] Jiayuan Zhang, Hu. Ziguo, Wei Ge, Yongjie Zhang, Tinghua Li, Jinghai Li, Application
Fig. 20. Evolution of the relative standard deviation (RSD) of concentration in a periodic of the discrete approach to the simulation of size segregation in granular chute flow,
chute flow. Ind. Eng. Chem. Res. 43 (18) (September 2004) 5521–5528.
T. Bhattacharya, J.J. McCarthy / Powder Technology 256 (2014) 126–139 139

[19] M. Barbolini, A. Biancardi, L. Natale, M. Pagliardi, A low cost system for the estima- [38] W.L. Vargas, J.J. McCarthy, Conductivity of granular media with stagnant interstitial
tion of concentration and velocity profiles in rapid dry granular flows, Cold Reg. fluids via thermal particle dynamics simulation, Int. J. Heat Mass Transf. 45 (2002)
Sci. Technol. 43 (1–2) (November 2005) 49–61. 4847–4856.
[20] O. Baran, D. Erta\cs, T.C. Halsey, G.S. Grest, J.B. Lechman, Velocity correlations in [39] Isabel Figueroa, Watson L. Vargas, Joseph J. McCarthy, Mixing and heat conduction
dense gravity-driven granular chute flow, Phys Rev E Stat Nonlin Soft Matter Phys in rotating tumblers, Chem. Eng. Sci. 65 (2) (2010) 1045–1054.
74 (5 Pt 1) (November 2006). [40] J.J. McCarthy, V. Jasti, M. Marinack, C.F. Higgs, Quantitative validation of the discrete
[21] D.V. Khakhar, J.J. McCarthy, J.M. Ottino, Radial segregation of granular mixtures in element method using an annular shear cell, Powder Technol. 203 (1) (2010)
rotating cylinders, Phys. Fluids 9 (1997) 3600–3614. 70–77.
[22] Deliang Shi, Adetola A. Abatan, Watson L. Vargas, J.J. McCarthy, Eliminating segrega- [41] C. Thornton, Force transmission in granular media, KONA 15 (1997) 81–90.
tion in free-surface flows of particles, Phys. Rev. Lett. 99 (2007) 148001. [42] R. Yang, R. Zou, K. Dong, X. An, A. Yu, Simulation of the packing of cohesive particles,
[23] Suman K. Hajra, Deliang Shi, J.J. McCarthy, Granular mixing and segregation in zig- Comput. Phys. Commun. 177 (1–2) (July 2007) 206–209.
zag chute flow, Phys. Rev. E. 86 (6) (December 2012) 061318. [43] Q.M. Tai, M.H. Sadd, A discrete element study of the relationship of fabric to wave
[24] J.M. Ottino, The Kinematics of Mixing: Stretching, Chaos, and Transport, Cambridge propagational behaviours in granular material, Int. J. Numer. Anal. Methods
University Press, New York, 1989. Geomech. 21 (1997) 295–311.
[25] J.J. McCarthy, T. Shinbrot, G. Metcalfe, J.E. Wolf, J.M. Ottino, Mixing of granular ma- [44] J.J. McCarthy, Micro-modeling of cohesive mixing processes, Powder Technol. 138
terials in slowly rotated containers, AIChE J 42 (1996) 3351–3363. (1) (2003) 63–67.
[26] C. Wightman, P.R. Mort, F.J. Muzzio, R.E. Riman, R.K. Gleason, The structure of mixtures [45] H. Li, J.J. McCarthy, Cohesive particle mixing and segregation under shear, Powder
of particles generated by time-dependent flows, Powder Technol. 84 (1995) 231–240. Technol. 164 (2006) 58–64.
[27] S.J. Fiedor, J.M. Ottino, Mixing and segregation of granular matter: multilobe forma- [46] Y. Tsuji, T. Kawaguchi, T. Tanaka, Discrete particle simulation of two-dimensional
tion in time-periodic flows, J. Fluid Mech. 533 (2005) 223–236. fluidized bed, Powder Technol. 77 (1993) 79–87.
[28] S.B. Savage, R. Dai, Studies of granular shear flows Wall slip velocities, ‘layering’ and [47] J.J. McCarthy, Turning the corner in segregation, Powder Technol. 192 (1) (2009)
self-diffusion, Mech. Mater. 16 (1–2) (1993) 225–238. 137–142.
[29] S.K. Hajra, D.V. Khakhar, Radial mixing of granular materials in a rotating cylinder: ex- [48] J. Schäfer, S. Dippel, E. Wolf, Force schemes in simulations of granular materials, J.
perimental determination of particle self-diffusivity, Phys. Fluids 17 (2005) 13101. Phys. I 67 (1991) 1751–1776.
[30] S.B. Savage, Disorder, diffusion and structure formation in granular flow, in: D. [49] J.J. McCarthy, J.M. Ottino, Particle dynamics simulation: a hybrid technique applied
Bideau, A. Hansen (Eds.), Disorder and Granular Media, Elsevier Science, to granular mixing, Powder Technol. 97 (1998) 91–99.
Amsterdam, 1993, pp. 255–285. [50] A. Di Renzo, Comparison of contact-force models for the simulation of collisions
[31] Tathagata Bhattacharya, Controlling size segregation of burden materials and pre- in DEM-based granular flow codes, Chem. Eng. Sci. 59 (3) (February 2004)
diction of void distribution in the upper part of blast furnace, LTP Project No: 525–541.
SI/LTP/02/05, Tata Steel, Research & Development Division, Tata Steel, Jamshedpur [51] H. Kruggel-Emden, E. Simsek, S. Rickelt, S. Wirtz, V. Scherer, Review and extension
831007, India, July 2007. of normal force models for the discrete element method, Powder Technol. 171 (3)
[32] P.A. Cundall, O.D.L. Strack, A discrete numerical model for granular assemblies, (February 2007) 157–173.
Geotechnique 29 (1979) 47–65. [52] H. Zhu, Z. Zhou, R. Yang, A. Yu, Discrete particle simulation of particulate systems:
[33] O. Walton, Application of molecular dynamics to macroscopic particles, Int. J. Eng. theoretical developments, Chem. Eng. Sci. 62 (13) (2007) 3378–3396.
Sci. 22 (1984) 1097–1107. [53] F.J. Massey, The Kolmogorov–Smirnov test for goodness of fit, J. Am. Stat. Assoc. 46
[34] K.L. Johnson, Contact Mechanics, Cambridge University Press, Cambridge, 1987. (253) (1951) 68–78.
[35] C. Thornton, K.K. Yin, M.J. Adams, Numerical simulation of the impact fracture and [54] N.W. Sharpe, J. Bridgwater, D.C. Stocker, Particle mixing by percolation, Trans. Inst.
fragmentation of agglomerates, J. Phys. D. Appl. Phys. 29 (1996) 424–435. Chem. Eng. 47 (1969) T114–T120.
[36] S.T. Nase, W.L. Vargas, A.A. Abatan, J.J. McCarthy, Discrete characterization tools for [55] M.H. Cooke, J. Bridgwater, A.M. Scott, Interparticle percolation: equipment de-
wet granular media, Powder Technol. 116 (2001) 214–223. velopment and mean percolation velocities, Trans. Inst. Chem. Eng. 56 (1978)
[37] K.Z.Y. Yen, T.K. Chaki, A dynamic simulation of particle rearrangement in powder 157.
packings with realistic interactions, J. Appl. Phys. 71 (1992) 3164–3173. [56] GSL—GNU Scientific Library, http://www.gnu.org/software/gsl/June 2010.

View publication stats

You might also like