Journal Pre-Proof: Powder Technology

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 64

Journal Pre-proof

Simulating a laboratory-scale cone crusher in DEM using


polyhedral particles

Flávio P. André, Luís Marcelo Tavares

PII: S0032-5910(20)30529-5
DOI: https://doi.org/10.1016/j.powtec.2020.06.016
Reference: PTEC 15510

To appear in: Powder Technology

Received date: 14 February 2020


Revised date: 4 May 2020
Accepted date: 5 June 2020

Please cite this article as: F.P. André and L.M. Tavares, Simulating a laboratory-scale cone
crusher in DEM using polyhedral particles, Powder Technology (2019), https://doi.org/
10.1016/j.powtec.2020.06.016

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2019 Published by Elsevier.


Journal Pre-proof

Simulating a laboratory-scale cone crusher in DEM using polyhedral

particles

Flávio P. André*, Luís Marcelo Tavares**

Department of Metallurgical and Materials Engineering, Universidade Federal do Rio

de Janeiro – COPPE/UFRJ, Cx. Postal 68505, CEP 21941-972, Rio de Janeiro, RJ

f
(Brazil)

oo
pr
Abstract
e-
Until now computer simulation has already successfully assisted in improving the
Pr

design and operation of cone crushers in the mining and aggregate industries. However,

simulations using the discrete element method, coupled to realistic breakage models, are
al

the next step towards designing confidently these machines in a fully virtual
rn

environment. The work shows the application of a novel breakage model fitted using
u

single-particle breakage information and polyhedral particles to simulate a laboratory-


Jo

scale cone crusher. Simulations of crushing three materials with widely different

characteristics were in good agreement with experiments for throughput and only

reasonably good for specific energy and product size. Sensitivity of simulations to

closed side setting, stroke and frequency showed good qualitative agreement to

experience. Simulations of crushing of blends containing mixtures of tough and soft

components then demonstrated that crusher throughput does not drop proportionally to

the increase in the amount of tough material in the feed.

1
Journal Pre-proof

Keywords: discrete element method, crushing, cone crusher, particle breakage, size

reduction

* Current address: SMI-JKMRC, The University of Queensland, Brisbane, Australia.

** Corresponding author: tavares@metalmat.ufrj.br

1. Introduction

Cone crushers are the most widely used type of crusher in the minerals and

f
oo
aggregate industries, given their robustness, ability to deal with rocks with a wide range

of abrasiveness, good energy efficiency as well as good control of product size [1].
pr
Creating a new cone crusher design, selecting a crusher model from a particular
e-
manufacturer, or optimizing a cone crusher that is already installed in a circuit has been
Pr

both an art and a science. Indeed, mathematical models such as the one proposed by W.

Whiten [2, 3] have been widely used in cone crusher optimization [4] but are of no
al

particular use in crusher design. On the other hand, the approach proposed by Evertsson
rn

[5, 6] has demonstrated to be powerful to aid cone crusher design as an alternative to

fully empirical models, but with still limited use in optimization of crushers in
u
Jo

operation. In spite of these developments and the nearly one century of experience with

these machines in industry, there are still opportunities for development, given the large

number of variables that influence crusher response as well as the wide variety of rock

types that are fed to it.

While successfully used in the minerals industry for nearly thirty years [7], the

discrete element method (DEM) [8] was first used only about 15 years ago [9] to

simulate cone crushing. This time lag may be explained by the fact that material flow

through the crushing chamber can only be realistically described if the rock undergoes

size reduction, thus requiring breakage models to be embedded in the DEM

2
Journal Pre-proof

environment [10]. As such, when coupled to models that can describe realistically

particle breakage, DEM can be a powerful tool to simulate cone crushers, since it allows

the user to investigate explicitly the effect of nearly all relevant design and operating

variables in crusher performance, such as chamber geometry, feeding arrangement, liner

design, speed, closed side setting, eccentric throw, feed rate, feed size and material.

Some of the attempts to simulate cone crushing in DEM were based on

approaches in which the primary fragments that make up the particle are resolved

throughout the entire simulation. Herbst and Potapov [9] used a model, called discrete

f
oo
grain breakage, in which triangular sub-particles interacted within each particle to

simulate a 2D crusher. Quist and Evertsson [11] and Johansson et al. [12] applied the
pr
bonded-particle model [13] to describe operation of cone crushers. In the model,
e-
clusters of spherical particles are bonded and, as bonds are progressively broken,
Pr

fragments are generated. Quist and Evertsson [11] simulated an industrial secondary

cone crusher operating under two closed side settings (CSS). They found relatively
al

good correspondence between simulation and experiments in respect to product size


rn

distribution, throughput and specific energy consumption, capturing qualitatively in the


u

simulations the high-frequency fluctuation of the pressure and power observed in


Jo

experiments. Johansson et al. [12] studied the application of the technique to a

laboratory cone crusher operating at two different speeds, with good agreement between

measured and simulated product size distributions, at least in the coarse size range.

However, they found difficulties in reaching choke-fed conditions in the simulations.

Both the discrete grain breakage and the bonded-particle model demanded large

computational effort, since even before being generated, fragments interact with each

other inside the parent particle. In the case of the bonded-particle model [11, 12] this

limitation led to the need to simulate only a quarter section of the crushers. Other

3
Journal Pre-proof

challenges in the method are volume conservation, since the clusters simulated are

artificially porous, and measuring the size distribution from the fragments [14].

An approach that is far more practical in simulating cone crushers is to use

particle replacement schemes, in which particles within a DEM simulation are replaced

instantaneously when a fracture criterion is met [15]. Given the convenience and lower

computational demand, particle replacement schemes involving spheres have been

widely used. Li et al. [16] simulated an industrial cone crusher by replacing each broken

particle by three spherical daughter particles. Their packing method allowed initial

f
oo
overlap of the fragments which replaced the parent particle when a fracture criterion

was met. In order to prevent the explosive decompression of the packed particles, they
pr
imposed a state in which only the fragments were initially allowed to move and large
e-
dissipation coefficients were used to remove the excess energy created by the overlap of
Pr

the fragments. After this relaxation, the main DEM simulation continued. They then

demonstrated that simulations on the effects of CSS and eccentric speed compared
al

favorably with the literature. Cleary and Sinnot [17] used a similar spherical particle
rn

replacement scheme to simulate several types of crushers, including a pilot-scale cone


u

crusher. They analyzed the power, energy dissipation and wear on the mantle for a
Jo

given set of conditions. However, given the important interaction between particles and

the crushing chamber in cone crushing, the use of spherical particles to represent both

parent and progeny particles lead to overly simplified descriptions of the process.

Lichter et al. [18] used a model based on replacement of particles by convex

polyhedral fragments to simulate a laboratory as well as an industrial cone crusher.

Such polyhedra have the advantage of more closely mimicking the fractal fragment

shapes found when breaking brittle material, besides conserving mass and volume. The

breakage model relied on the energy-specific breakage rate and breakage distribution

4
Journal Pre-proof

functions, which were originally proposed for ball milling [19]. The authors showed

very good agreement between measured and simulated product size distributions as well

as throughputs for two crushing tests. No information on power draw was provided.

Delaney et al. [20] and Cleary et al. [21] used super-quadric particles on a

geometric packing replacement scheme involving also non-round fragments to simulate

an industrial cone crusher. They combined simulation of the resolved fragments

produced by particle replacement to the unresolved fines, which are later added at a

post-processing stage, thus mimicking the fines which would not be practically

f
oo
simulated in DEM. This allowed the prediction of product size distributions down to

very fine sizes, limited only by the fineness of the fragments measured in the breakage
pr
characterization, which relied on JK Drop Weight Tests [3]. In their studies, the authors
e-
analyzed in great detail the intricacies of the flow of material through the crushing
Pr

chamber. Also, in the most comprehensive parametric study of cone crushers using

DEM to date, Cleary et al. [21] studied the effect of CSS, eccentric speed, friction factor
al

and rock strength on crusher throughput, power and product size distribution.
rn

In spite of the relative success in the past in application of DEM to cone


u

crushing there are still opportunities to improve upon the earlier approaches by
Jo

describing more realistically the response of particles to the stressing environment. This

can be done, for instance, by taking into account more accurately the size-dependent

breakage probabilities, the weakening by repeated stressing as well as the stressing-

energy dependent progeny size distribution. The present work takes advantage of a

detailed breakage model recently implemented in a commercial DEM simulation

platform [22] and calibrated on the basis of single-particle breakage data to analyze the

effect of closed side setting, crusher speed, eccentric throw and feed blend on crusher

throughput, power, specific energy and product size distribution.

5
Journal Pre-proof

2. Particle breakage modeling background

The breakage model implemented in version 4.3 of Rocky DEM describes body

breakage of convex polyhedral particles when subjected to stresses of varied

magnitudes. This model takes into account the energy dissipated in the contact between

two elements to decide whether the particle will break or not. The model itself needed to

be adapted from a continuous to a discrete approach, where each particle behaves

independently according to the stresses they are subject to. Few modifications in the

f
oo
model equations were necessary to adapt it to the discrete environment given by DEM

[22]. Regarding particles fracture energy, each new particle entering the simulation is
pr
assigned, by Monte Carlo simulation, a random value of specific fracture energy (𝐸)
e-
following an upper-truncated lognormal distribution [23, 24] based on the breakage
Pr

parameters that describe the material, according to


al

1 ln 𝐸 ∗ − 𝑙𝑛 𝐸50
𝑃𝑜 (𝐸) = [1 + erf ( )] (1)
2
rn

√2𝜎 2

and
u

𝐸𝑚𝑎𝑥 𝐸
Jo

𝐸∗ = (2)
𝐸𝑚𝑎𝑥 − 𝐸

where 𝐸 is the specific fracture energy of the particle, 𝐸𝑚𝑎𝑥 is the upper truncation

value of the distribution, 𝐸50 and 𝜎² are the median and the geometric variance of the

distribution, respectively.

The relationship between particle size and median fracture energy, hereby called

size function, is described using the expression

6
Journal Pre-proof

𝜑
𝑑𝑜
𝐸50 = 𝐸∞ [1 + ( ) ] (3)
𝑑

where 𝐸∞ , 𝑑𝑜 and 𝜑 are model parameters that must be fitted to experimental data and

𝑑 is the size of the particle.

Rocky DEM considers that in a contact between particles or between a particle

and a liner of a crusher, the energy dissipated in the contact is split between the bodies,

according to the stiffness of the surfaces in contact [25], as well as the particle volume.

f
oo
As such, from the values of stiffness and sizes of the bodies in contact, the specific

energy actually dissipated in the particle during stressing is given by Ek .


pr
When the first contact that does not lead to breakage occurs, particles will
e-
sustain damage and the fracture energy of each particle will be reduced according to the

expression [24]
Pr
al

𝐸𝑛𝑒𝑤 = 𝐸𝑜𝑙𝑑 (1 − 𝐷) (4)


rn

where 𝐸𝑛𝑒𝑤 and 𝐸𝑜𝑙𝑑 are the fracture energies of the particle after and before the
u
Jo

collision event, respectively. D is the damage sustained by a particle after a contact that

does not promote breakage and is given by [26]

2𝛾 (5)
2𝛾 𝐸𝑘 5
𝐷=⌈ ⌉
(2𝛾 − 5𝐷 + 5) 𝐸

where 𝛾 is the damage accumulation coefficient, which characterizes the amenability of

a material to sustain damage prior to catastrophically breaking.

7
Journal Pre-proof

When the specific energy dissipated in the particle (𝐸𝑘 ) due to the contact

between particles and/or particles and the machine geometry surpasses the specific

fracture energy of the particle (𝐸), breakage will occur and the fragments will be

generated based on the Laguerre-Vonoroi algorithm. Progeny fineness will be described

on the basis of the value of 𝑡10, corresponding to the percentage in weight of the parent

particle that passes through a sieve with aperture of 10% of the original particle size [22,

26], so that

f
oo
𝑡10 = 𝐴 [1 − exp(−𝑏′ ) ] (6)

pr
where 𝐴 and 𝑏′ are model parameters fitted to experimental data, in which 𝐴
e-
corresponds to the maximum value of 𝑡10 that can be achieved when breaking a material
Pr

in a single stressing event. From the value of t10 the complete particle size distribution

can be expressed with the aid of the incomplete beta function, given by [27]
al
rn

𝑡10 ⁄100 (7)


100
𝑡𝑛 (𝑡10 ) = 1 ∫ 𝑥 𝛼𝑛−1 (1 − 𝑥 )𝛽𝑛−1 𝑑𝑥
u

∫0 𝑥 𝛼𝑛−1 (1 − 𝑥 )𝛽𝑛−1 𝑑𝑥 0
Jo

where the cumulative percentage in mass of the particles passing a screen (𝑡𝑛 ) with

relative size 𝑥 is calculated by the given 𝑡10 value of the distribution. The fragments of

the original particle will then be assigned a particle fracture energy based on the

distribution given by Equations (1) and (2), according to their particle size.

The remaining energy of the contact will be used to promote breakage of the

fragments originated on the first breakage event until its full dissipation. It is possible to

observe, based on Equation (6), that every breakage event involving particles of the

8
Journal Pre-proof

same material can be normalized into an identical progeny size distribution. As such,

the model as adapted and implemented in Rock DEM uses the concept of primary

normalized breakage function, which defines a constant material-specific size

distribution resulting from each breakage event, as well as the concept of describing a

single impact event as successive primary breakage events, as suggested by Saeidi et al.

[28].

The size distribution of the progeny particles resulting from a breakage event is

represented by the Laguerre-Voronoi tessellation [29] based on Equations (6) and (7).

f
oo
The application of the Voronoi diagram in the Laguerre geometry provides appropriate

control over the size of the fragments based on a densely packed set of spheres of a pre-
pr
determined size distribution, thus allowing the generation of cells that are similar to the
e-
sphere size distribution. The Laguerre-Voronoi tessellation can be expressed
Pr

mathematically as
al

{𝐶𝑃𝑖 } = {𝑥 ∈ 𝑆 3 ∶ ‖𝑃𝑖 − 𝑥 ‖ − 𝑟𝑖2 ≤ ‖𝑃𝑗 − 𝑥 ‖ − 𝑟𝑗 ²} and 𝑗 = 1, 2, … , 𝑁 ∶


(8)
rn

𝑗≠𝑖
u
Jo

where 𝑆 3 is the space, 𝑥 is the position of a generic point in 𝑆 3 , 𝑃𝑗 is the position of the

j-th nucleus, 𝐶𝑃𝑖 and 𝑟𝑖2 are the cell and the weight associated to the nucleus 𝑃𝑖 ,

respectively.

The size of each particle is determined in Rocky DEM by the largest size of the

particle and its perpendicular size. Particle size is based upon the dimensions of a square

hole that is just large enough for the particle to pass through.

9
Journal Pre-proof

The model has been verified and validated in great detail for several materials on

the basis of data from single-particle and unconfined particle-bed impact breakage tests

[22].

3. Experimental

3.1. Materials

Three materials have been selected for this study, namely a limestone, a

granulite and a copper ore. These materials were chosen for their different breakage

f
oo
responses, thus allowing to test the model over a wide range of breakage strengths. A

summary of physical characteristics of the samples is given in Table 1, including results


pr
from standard crushability and grindability tests [3, 30], for reference. Bond ball mill
e-
work index values reported were obtained with a closing sieve of 150 µm, whereas the
Pr

standard Drop Weight Test (JKDWT) A*b value was estimated on the basis of particles

contained in the size range from 63 to 13.2 mm.


al
rn

TABLE 1
u
Jo

3.2. Crushing tests

Crushing experiments were carried out in a laboratory scale short-head cone

crusher (Denver No. 12) (which was lately modeled using the computer aided design

tool AutoCAD® as presented in Figure 1). The crusher consists mainly of a mantle and a

concave cloistered by a housing structure. A feed hopper is located at the top of it to

allow handling of larger quantities of sample. The crusher presents a feed opening of 29

mm, a stroke of 0.45 mm and a mantle of 328 mm in diameter at the bottom with a

length of 155 mm and a side angle of 46% in relation to horizontal. The axis of the

10
Journal Pre-proof

mantle is inclined in relation to the vertical and the tilted axis coincides at a pivot point

located below the base of the mantle. The concave has a height of 120 mm and internal

diameters on the top and on the bottom of 225 and 323 mm, respectively. The closed

side setting (CSS) of the crusher used in the experiments was 5.0 mm. The motor runs

at about 1725 rpm, transferring torque via a rubber belt to the crusher bowl that turns in

a precession motion at a frequency of 616 ±5 rpm. The main details of the crusher are

presented in an annotated cutaway view in Figure 2.

f
oo
FIGURE 1

FIGURE 2
pr
e-
Experiments consisted of manual feeding samples containing from 10 to 15 kg
Pr

of material in the size range of 22.4-16.0 mm to the crusher and collecting timed

samples of the product. Choke-fed conditions were maintained during the tests. Material
al

generated in the beginning and final portions of the test, when non-choke conditions
rn

prevailed, was discarded. Size analyzes of the crusher discharge were measured by wet-
u

dry sieving. From the sampling time and the sample weights, the throughput for each
Jo

test was estimated. Timed samples for size analyses and for estimating throughput were

taken in duplicate and average values were used. Crusher power was measured during

the tests and the average power during collection of the sample was measured. The net

power was then calculated by subtracting the crusher power in operation from the no-

load power measured when the crusher was idling.

4. Simulation

4.1. Simulation setup

11
Journal Pre-proof

Simulations were run in a developer version of the software Rocky DEM, which

was later released as Rocky DEM version 4.3. The hardware configuration used in the

present work consisted of a workstation with 32 GB DDR4 3200 MHz of RAM, using a

Nvidia Titan V GPU card and an Intel ® Core™ i7-8700K CPU @ 3.70 GHz processor.

The average simulation pace for all the scenarios tested was about 4 hours per second of

crusher operation.

In the crusher, the motion consisted of the cone tilting at a certain angle rotating

around a vertical axis. In order to describe it the software requires information of the

f
oo
coordinates of the pivot point from the origin and the rotational velocity. These are the

angular velocity around the vertical axis and the initial orientation of the cone based on

the sine and cosine of its inclined axis.


pr
e-
In Rocky DEM the cone crusher motion applied to a geometry is described as a
Pr

free body rotation motion, which means that the cone will perform its fixed precession

movement around the vertical axis but it is free to spin around its own inclined axis in
al

response to outside forces, such as forces originated from particle contacts. This free
rn

rotation movement around the axis is then prevented by the torque imposed by particles
u

when in contact with the mantle, describing accurately the real movement of a cone
Jo

crusher.

The information required by Rocky DEM to represent the motion of the cone in

simulations for this particular crusher was not readily available, being necessary to

perform measurement of the operating conditions of the crusher to collect the

appropriate data. Since the movement of the cone is repeated several times during each

simulation, modeling of the crushing operation was carefully examined in order to avoid

accumulating errors of any kind, which could lead to biased results in the simulations.

12
Journal Pre-proof

The eccentric throw of the cone in relation to the horizontal was measured using

play dough and its inclination was achieved using a clinometer. The values of 1.5 mm

and 0.5º were obtained, respectively. The pivot point, which corresponds to the point in

which the shaft of the cone will pivot around, was determined using trigonometry and it

is positioned 171.8 mm below the base of the cone in the vertical axis. The validation of

the measurements was demonstrated by comparing the expected values of the closed

and open side setting with animations provided by the motion preview tool available in

Rocky DEM, as presented in Figure 3.

f
oo
FIGURE 3
pr
e-
Besides describing the correct precession movement of the cone, its physical
Pr

properties are necessary so the rotation motion of the cone around its own inclined axis

can be represented appropriately. The user must input the mass of the cone, its center of
al

gravity in relation to the origin and the moment of inertia for each axis. Default
rn

parameters available in Rocky DEM, which are meant to be used for industrial-scale
u

crushers, are likely to overestimate the moment of inertia of a laboratory or pilot scale
Jo

crusher, preventing the forces on the mantle from being balanced and reaching a steady-

state condition. The volume of the cone and its center of gravity were determined using

the software MeshLab whereas the mass of the cone was estimated based on its volume

and considering it as a solid geometry made of steel. The moment of inertia for the

vertical axis was estimated based on the radius of the cone equal to 163.9 mm. Table 2

shows the mass/volume parameters of the cone geometry.

TABLE 2

13
Journal Pre-proof

Simulations were initially conducted by setting up the crusher to the

experimental values of closed side setting, eccentric throw and crusher speed and using

a feed containing a desired size distribution to mimic the conditions found in

experiments. Choke-feed conditions were achieved by filling the feeder of the crusher

before starting its operation and by adopting a feed flow rate higher than the expected

throughput, allowing the filling of the feed hopper during the simulations. Different

feed flow rates were used according to the material tested to offset the differences found

f
oo
in throughput for each material in experiments. Feed flow rates of 2.0, 2.5 and 3.0 t/h

were used for simulations of copper ore, granulite and limestone, respectively. As in
pr
any cone crusher operating under choke-fed conditions, the discharge flow rate is
e-
defined by flow characteristics of the material through the crushing chamber.
Pr

Simulations were assessed after steady-state conditions were reached on the basis of

discharge flow rate, particle size distribution of product, power draw and specific power
al

consumption.
rn

On a second stage of the work, simulations were carried out varying the
u

operating conditions of the crusher, such as the CSS for the three materials and the
Jo

speed and eccentric throw for the copper ore. In a final stage of the study, simulations of

the crusher with multi-component blends containing different proportions (in weight) of

tough rock (copper ore) and soft rock (limestone) were performed and the results were

analyzed.

4.2. Material and contact parameters

14
Journal Pre-proof

The contact models adopted in all simulations were the hysteretic linear spring

model for the normal component of the force and the linear spring Coulomb limit for

the tangential component of the force [31, 32].

Contact detection using polyhedral particles in Rocky DEM takes into

consideration relevant geometrical parameters of particles whenever they are detected

within neighboring distance from other particles or boundaries. Because of their shape,

contact between polyhedra might occur between vertices, edges and faces. The

calculation of the forces arising from the contacts accounts for all the possible

f
oo
interaction scenarios to determine the normal contact force based on the normal overlap

between elements [31]. Further details of the contact models can be found elsewhere

[22].
pr
e-
Table 3 presents the contact parameters adopted, which were obtained on the
Pr

basis of bench-scale testing of the copper ore [14]. Since this work is focused on the use

of the breakage model in cone crusher simulations and in performing its initial
al

validation, these parameters were also used to describe flow of granulite and limestone,
rn

in spite of their different characteristics. In practice, proper calibration of contact


u

parameters is critical prior to DEM simulations, and the authors acknowledge that.
Jo

However, by using the same contact parameters for the different materials the authors

wish to demonstrate the dominating influence of breakage parameters rather than

contact parameters in DEM simulation of a flow-restricted machine such as a cone

crusher.

TABLE 3

15
Journal Pre-proof

The geometry of the particles was chosen in order to mimic the shape of a real

particle. However, it was observed that the use of only one particle shape promoted the

appearance of unrealistic voids among the particles, affecting particle flow inside the

crushing chamber. Therefore, four different particle shapes, ranging from a nearly

spherical shape to a very angular shape, were used in simulations in order to reach an

appropriate packing of the material. Table 4 summarizes the input parameters for the

different particle shapes adopted.

f
oo
TABLE 4

pr
The value of Bulk Young’s modulus of 1x109 N/m² was set for steel, whereas
e-
the values for the different rock types were set to 1x108 N/m². Being the Young’s
Pr

modulus of the particles 1/10th of that of steel, a proper split of the energy between the

elements involved in the contacts became possible. These values are in the order of
al

magnitude of those for the materials since, as correctly pointed out by Delaney et al.
rn

[20], the stiffness of the rock in a crusher simulation should be fairly realistic of that of
u

the material and should not be artificially reduced to speed-up a simulation, given the
Jo

very high stresses that are present in the system.

4.3. Breakage model parameters

Parameters in the model equations described in section 2 fitted from single-

particle impact-breakage experiments using impact load cell devices [23, 33] were

collected from previous publications [34, 35] and values are presented in Table 5. The

16
Journal Pre-proof

value of the upper truncation parameter of the lognormal distribution (Emax) (Equation

2) was set to four times the value of E50 for all materials.

TABLE 5

The minimum global particle size assigned in all of the simulations

corresponded to 1/10th of the representative size (1.9 mm). Although relatively coarse,

f
this value allowed conducting simulations until steady-state conditions were reached in

oo
the order of 30-40 hours, which made exploring a variety of operating conditions

pr
possible. In each test the average mass inside the crushing chamber was about 3.3 kg.
e-
5. Results and discussion
Pr

5.1. Simulation validation

Figure 4 shows a snapshot of a crushing simulation of granulite particles in the


al

laboratory cone crusher. It is possible to notice that the simulation presents adequate
rn

packing of material inside the crusher chamber and that the level of material in the feeder
u

of the crusher indicates that the crusher is operating under choke-fed conditions.
Jo

FIGURE 4

Figure 5 illustrates the dynamics of the throughput of the crusher receiving

different feed materials. As simulations involved three distinguishable materials

according to their breakage strength, varying from the tough copper ore to the soft

limestone, it becomes clear that the model was able to capture the effect of the different

breakage characteristics of the materials on the throughput. The first two seconds in the

17
Journal Pre-proof

simulations are only dedicated to fill the feeder of the crusher, so that the movement of

the cone only starts when this time has elapsed. Analyzes of the throughput account for

the moving average flow detected within each second. The three simulations tended to

reach steady-state conditions between five to eight seconds of simulation (which

corresponds to three to six seconds of crusher operation). The crushing simulation of the

softer material (limestone) reached steady-state conditions in about three seconds of

crusher operation, whereas nearly six seconds were required to reach steady-state

conditions in the case of the tougher copper ore. Differences between the flow rates of

f
oo
materials being processed by the crusher under steady-state conditions are quite clear in

Figure 5 (top), indicating that the model described well the influence of material
pr
strength in crusher throughput. Indeed, tougher materials take longer within the
e-
crushing chamber to become fine enough to move towards the crusher discharge, thus
Pr

leading to a reduction in throughput and to a longer time required to reach steady-state

conditions.
al

The power draw in simulations was also investigated. This value corresponds to
rn

the power required to move the geometry at a specified speed given the stresses applied
u

by the particles to the surface of the geometry. While soft materials, such as limestone,
Jo

tend to break at lower energy, imposing less resistance to the movement of the cone,

copper ore particles will break at higher energies, demanding more power from the

crusher to perform its precession motion. This becomes evident in Figure 5 (bottom),

which compares the power draw for the three materials simulated. It is possible to

observe that the power draw stabilizes at three different levels according to the material

being simulated, demanding more power in crushing simulations of copper ore and less

power in simulations of limestone, as expected. It is important to notice in Figure 5

(bottom) that during the first second and a half of crusher operation (2 to about 3.5

18
Journal Pre-proof

seconds of simulation), simulations returned values of power draw that were higher than

the readings from the rest of the simulation. These peaks are a consequence of the free

body rotation motion associated to the cone, allowing it to spin freely around its own

axis until reaching an equilibrium condition in which the forces acting on its surface

prevent it from rotating, as observed in reality. Beyond that, Fluctuations in power draw

are associated to individual compression events in the crushing chamber, as observed in

the literature [11]. Peaks in power draw, such as those observed in Figure 5 (bottom),

can impose a limit on a crusher operation when dealing with a very tough ore,

f
oo
demanding a higher installed power of the motor.

The comparison between simulated and experimental results for the three
pr
materials has also been performed and it is presented in Table 6. Simulation results
e-
presented good agreement to experiments. An average of 20% more throughput was
Pr

detected in simulations when compared to experimental data. Moreover, the simulated

power draw and the experimental net power measured using a power meter attached to
al

the cone crusher presented only reasonable agreement. The lower values of simulated
rn

power draw, in addition to the overestimation of throughput, resulted in


u

underestimations in the specific energy consumption of the simulations. Nevertheless,


Jo

lower values of power draw in simulations were not surprising since they do not take

into account the existing power dissipation in experimental tests, which can be

particularly large for a machine of this size with a transmission that relies on a rubber

belt [36]. It is also important to bear in mind that breakage parameters were not fitted on

the basis of the crushing experiments, having been estimated previously on the basis of

single-particle breakage experiments. In addition to that, the same contact parameters

were used for all materials and an opportunity exists in modifying them so that a better

match between experiments and simulations can be reached. However, the fact that the

19
Journal Pre-proof

contact parameters used in all simulation were from the copper ore (Table 3) and that

the deviations found between the performance values for this material were not smaller

than for the other two, reinforces the points raised above.

TABLE 6

Figure 6 presents the comparison between simulated and experimental particle

size distributions for crushing experiments of the three materials. It is evident that

f
oo
simulations were able to identify that limestone presented a finer size distribution in the

product than both copper ore and granulite. It is also evident that coarser particles were
pr
predicted in the crusher product than observed in the experiments. Such deviation on the
e-
coarse portion of the curves may be explained by the tendency of the Laguerre-Voronoi
Pr

tessellation scheme in generating flaky fragments on a breakage event, preventing them

from rebreakage as they move downwards in the crushing chamber towards the
al

discharge. Figure 4 illustrates particles coarser than 9 mm (represented in red), which


rn

match the top size found in experiments, leaving the crusher in the simulations owing to
u

their flakyness.
Jo

On the fine end of the curves in Figure 6 there was no surprise on the inability of

the simulation to match the experimental data, which was due to the minimum size of

only 1/10th of the representative size of the feed. The inclusion of unresolved fines in a

post-processing stage using the model described in section 2 following the approach

adopted by Delaney et al. [20] and Cleary et al. [21] would be an alternative to

circumvent this limitation. However, such feature was not yet available in the developer

version of Rocky DEM used in the present work.

20
Journal Pre-proof

FIGURE 6

5.2. Effect of closed side setting

Results from simulations by varying the closed side setting (CSS) from 4.0 to

8.6 mm are summarized in Figure 7 (top), which shows the well-known effect of

increasing throughput with CSS [37]. Both the slope and the vertical position of the

lines, however, clearly varied with material.

f
oo
FIGURE 7

pr
The power draw from the crusher, omitted for brevity, dropped only modestly
e-
with the increase in CSS, so that the specific energy consumption reduced significantly,
Pr

as shows Figure 7 (bottom). This is also qualitatively consistent with data from the

literature [37], as well as crusher manufacturers catalogs. As expected, simulations


al

demonstrated that increasing the CSS results in a coarser product at the discharge of the
rn

crusher. Results for the copper ore indicate that the P80 varied almost linearly from 6.5
u

mm for the CSS of 4.0 mm to 11.7 mm for the CSS of 8.6 mm.
Jo

5.3. Effect of eccentric speed

Although not common, crusher speed can be varied in a cone crusher to optimize

its performance [38]. In the simulations of size reduction of copper ore, the crusher

speed was varied from 308 rpm to 1232 rpm. Figure 8 shows that throughput initially

increases with the increase in cone frequency, reaches a maximum and then drops for

higher velocities. This has been demonstrated in the past using analytical models [5, 38]

and characterizes a transition from subcritical to supercritical speed. While at low

21
Journal Pre-proof

speeds particles are able to slide and fall freely until they are captured when the crusher

starts to nip, at supercritical speeds particles are nipped while still moving downwards

in the crusher chamber, thus restricting the material flow through the crusher.

In addition, the crusher power (omitted for brevity) increases as the eccentric

speed increases. The result is that the specific energy increases initially modestly with

the increase in speed, but then significantly as the throughput starts to decrease at higher

speeds, identifying the inefficiency associated to operating under this condition for the

crusher in question (Figure 8). The fineness of the product also increases with the

f
oo
specific energy until 924 rpm, beyond which the product becomes also coarser, which

further demonstrates the low attractiveness of operating at such high speeds.


pr
e-
FIGURE 8
Pr

5.4. Effect of eccentric throw

The eccentric throw also plays an important role in cone crusher throughput. In
al

order to study the effect of stroke, the cone geometry had its inclination varied. The
rn

CSS was maintained constant at 5.0 mm throughout the simulations by displacing the
u

cone vertically. An increase in 0.25° in the inclination of the cone changed the stroke by
Jo

0.25 mm. Simulations with copper ore were carried out by varying the stroke from 0.45

mm to 0.95 mm. The effect of different strokes on the throughput as well as the power

draw of the crusher is presented in Figure 9. It shows that, within the range of values

studied, it increased positively the crusher throughput.

The power draw was strongly influenced by the changes on the eccentric throw,

resulting in larger demand of power when operating with a larger stroke. Still, its rate of

growth was not as great as the one observed for throughput, resulting in a reduction in

22
Journal Pre-proof

the specific energy when increasing the size of stroke. The influence of it in the specific

energy is also presented in Figure 9.

FIGURE 9

Nevertheless, changing the size of the stroke did not affect significantly the size

distribution of the product, although the product became marginally finer when

operating at larger throws. This is mainly due to the rebreakage of fragments in the

f
oo
crushing chamber as larger strokes are able to nip more particles over time. A

comparison between the size distributions of the products can be observed in Figure 10.
pr
The P80 of the product varied from 7.7 mm for a stroke of 0.45 mm to 7.3 mm for a
e-
stroke of 0.95 mm, thus demonstrating a potential benefit for using larger strokes. In
Pr

practice, however, the maximum throw that may be used for a particular rock and

crusher is related to the stiffness of the feed material and the resulting stresses that may
al

result from operating at the selected throw, which may reach critical stresses that can be
rn

stood by parts of the machine.


u
Jo

FIGURE 10

5.5. Effect of feed blend

In both the minerals and aggregate industries crushers are fed with whatever

rocks types that make up the mineral deposit. Given the significant effect of rock

properties that were observed on crusher performance (section 5.1), it is worthwhile

investigating the effect of blends on the feed. Mixtures containing copper ore (tougher

component) and limestone (softer component) at different proportions in weight in the

23
Journal Pre-proof

feed were simulated to analyze the effect of blending of the crusher feed. Figure 11

illustrates its effect in the crusher throughput. As expected, the increase in the

proportion of copper ore in the feed was responsible for reducing the throughput due to

the higher toughness of the copper ore. However, this reduction in throughput was not

proportional to the increase in the content of copper ore in the mixture. In fact, it shows

that a relatively small proportion of copper ore (25%) reduced in 35% the throughput of

the crusher. No study was found in the literature dealing with mixtures of components

in the crusher feed. However, this result is consistent with those from studies on the

f
oo
impact of mixtures on the performance of ball [39] and semi-autogenous mills [40].

The increase in the proportion of tough material in the feed also impacted
pr
directly the power consumption of the crusher. Increasing the proportion of copper ore
e-
in the feed blend reduced the throughput and increased the power draw, therefore, the
Pr

specific energy of the crusher increased significantly. This is also shown in Figure 11,

which demonstrates that specific energy increased proportionally to the content of tough
al

material in the feed.


u rn

FIGURE 11
Jo

The proportion of each material at the discharge of the crusher for different

blends was equivalent to the proportion being fed to the crusher, indicating that the soft

limestone was retained by copper ore particles inside the crusher. This effect becomes

clear when analyzing the size distribution of the product for each blend. When

increasing the proportion of limestone being fed to the crusher, the product becomes

finer, as presented in Figure 12, indicating the occurrence of rebreakage of limestone

particles retained in the crusher chamber.

24
Journal Pre-proof

FIGURE 12

6. Conclusions

A laboratory-scale cone crusher has been modeled and successfully simulated

using the discrete element method and a detailed breakage model involving polyhedral

particles in Rocky DEM. Comparisons between measured and predicted throughput for

the crusher in size reduction of three widely different materials showed good agreement,

f
oo
while only reasonable agreement has been observed between measured and predicted

product size distributions and specific energy consumption.


pr
A sensitivity analysis of the crusher performance to a number of operating
e-
variables demonstrates trends which are in qualitative agreement with the literature and
Pr

industrial practice, with throughput increasing and specific energy decreasing with

closed side setting; throughput presenting a maximum for a particular crusher


al

frequency; whereas throughput and product fineness increasing with crusher throw.
rn

The investigation of the laboratory crusher performance when dealing with feed
u

blends has shown that throughput drops more significantly than the increase in the
Jo

proportion of tough material in the feed, whereas specific energy increases

approximately linearly with it.

In summary, the work shows that the combination of a breakage model that

captures the variability in the fracture energies of the feed, weakening whenever they do

not break, and a material-specific breakage distribution, coupled to the description of a

particle replacement scheme with polyhedra resulted in physically-sound simulations of

a cone crusher.

25
Journal Pre-proof

Acknowledgements

The authors would like to thank the partial financial support from ESSS and ME

Elecmetal to this investigation. We also thank the additional financial support by the

Brazilian research agencies CNPq (grant number 310293/2017-0) and FAPERJ (grant

number E-26/202.574/2019). The suggestions by Dr. Alexander Potapov and Prof.

Rodrigo M. de Carvalho during the course of the work are deeply appreciated.

References

f
oo
1. B.A. Wills, T.J. Napier-Munn, Mineral Processing Technology, Butterworth-

Heinemann, 7th edn., 2007.


pr
2. S. Morrell, T.J. Napier-Munn, J. Andersen, The prediction of power draw in
e-
comminution machines. In: Comminution: Theory and Practice, SME, Littleton,
Pr

pp. 233–247, 1992.

3. T.J. Napier-Munn, S. Morrel, R.D. Morrison, T. Kojovic, Mineral Comminution


al

Circuits: Their Operation and Optimization, JKMRC Monograph Series, 1996.


rn

4. R.P. King, Simulation — the modern cost-effective way to solve crusher circuit
u

processing problems, Int. J. Miner. Process. 29 (1990) 249-265.


Jo

5. C.M. Evertsson, Output prediction of cone crushers, Miner. Eng. 11 (1998) 215-

231.

6. C.M. Evertsson, Modeling of flow in cone crushers, Miner. Eng. 12 (1999)

1479-1499.

7. B.K. Mishra, R.K. Rajamani, The discrete element method for the simulation of

ball mills, Appl. Math. Model. 16 (1992) 598-604.

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

assemblies. Géotechn. 29 (1979) 47–65.

26
Journal Pre-proof

9. J.A. Herbst, A.V. Potapov, Making a discrete grain breakage model practical for

comminution equipment performance simulation, Powder Technol. 143-144

(2004) 144-150.

10. N.S. Weerasekara, M.S. Powell, P.W. Cleary, L.M. Tavares, M. Evertsson, R.D.

Morrison, J. Quist, R.M. Carvalho, The contribution of DEM to the science of

comminution, Powder Technol. 248 (2013) 3–24.

11. J. Quist, C.M. Evertsson, Cone crusher modelling and simulation using DEM,

Miner. Eng. 85 (2016) 92–105.

f
oo
12. M. Johansson, J. Quist, M. Evertsson, E. Hulthén, Cone crusher performance

evaluation using DEM simulations and laboratory experiments for model


pr
validation, Miner. Eng. 103-104 (2017) 93-101.
e-
13. D.O. Potyondy, P.A. Cundall, A bonded-particle model for rock, Int. J. Rock
Pr

Mech. Min. Sc. 41 (2004) 1329-1364.

14. N. Jiménez-Herrera, G.K.P. Barrios, L.M. Tavares, Comparison of breakage


al

models in DEM in simulating impact on particle beds, Adv. Powder Technol. 29


rn

(2018) 692–706.
u

15. P.W. Cleary, Recent advances in DEM modelling of tumbling mills, Miner.
Jo

Eng. 14 (2001) 1295-1319.

16. H. Li, G. McDowell, I. Lowndes, Discrete element modelling of a rock cone

crusher, Powder Technol. 263 (2014) 151-158.

17. P.W. Cleary, M.D. Sinnott, Simulation of particle flows and breakage in

crushers using DEM: Part 1 - Compression crushers. Miner. Eng. 74 (2015)

178–197.

18. J. Lichter, K. Lim, A. Potapov, D. Kaja, New developments in cone crusher

performance optimization, Miner. Eng. 22 (2009) 613-617.

27
Journal Pre-proof

19. J.A. Herbst, D.W. Fuerstenau, Scale-up procedure for continuous grinding mill

design using population balance models, Int. J. Miner. Process. 7 (1980) 1-31.

20. G.W. Delaney, R.D. Morrison, M.D. Sinnott, S. Cummings, P.W. Cleary, DEM

modelling of non-spherical particle breakage and flow in an industrial scale cone

crusher, Miner. Eng. 74 (2015) 112–122.

21. P.W. Cleary, M.D. Sinnott, R.D. Morrison, S. Cummings, G.W. Delaney,

Analysis of cone crusher performance with changes in material properties and

operating conditions using DEM. Miner. Eng. 100 (2017) 49–70.

f
oo
22. L.M. Tavares, F.P. André, A. Potapov, C. Maliska Jr., Adapting a breakage model

to discrete elements using polyhedral particles, Powder Technol. 362 (2020) 208-

220.
pr
e-
23. L.M. Tavares, R.P. King, Single-particle fracture under impact loading, Int. J.
Pr

Miner. Process. 54 (1998) 1-28.

24. L.M. Tavares, R.P. King, Modeling of particle fracture by repeated impacts
al

using continuum damage mechanics, Powder Technol. 123 (2002) 138-146.


rn

25. L.M. Tavares, R.M. Carvalho, Modeling ore degradation during handling using
u

continuum damage mechanics, Int. J. Miner. Process. 101 (2011) 21-27.


Jo

26. L.M. Tavares, Analysis of particle fracture by repeated loading as damage

accumulation, Powder Technol. 190 (2009) 327-339.

27. R.M. Carvalho, A. Secchi, L.M. Tavares, A new energy-based breakage

function model and optimization procedure for mechanistic comminution

models. In: 14th European Symposium on Comminution and Classification,

2015, Gothenburg, pp. 1-4.

28. F. Saeidi, L.M. Tavares, M. Yahyaei, M. Powell, A phenomenological model of

single particle breakage as a multi-stage process, Miner. Eng. 98, 90–100, 2016.

28
Journal Pre-proof

29. H. Imai, M. Iri, K. Murota, Voronoi diagram in the Laguerre geometry and its

applications, SIAM Journal on Computing, 14 (1985), 93-105.

30. L.M. Tavares, M.A.W. da Silveira, Comparison of measures of rock

crushability. In: Fine Particle Technology and Characterization (M. Yekeler,

Ed.), Research Signpost, 2008, pp. 1-20.

31. ROCKY DEM, DEM Technical Manual 4.2, 2018.

32. O.R. Walton, R.L. Braun, Stress calculations for assemblies of inelastic spheres

in uniform shear, Acta Mech. 63 (1986) 73-86.

f
oo
33. L.M. Tavares, Breakage of single particles: quasi-static. In: Handbook of

Particle Breakage (A.D. Salman, M. Ghadiri, M.J. Hounslow, Eds.), Elsevier,

vol. 12, pp. 3-68, 2007.


pr
e-
34. L.M. Carvalho, L.M. Tavares, Predicting the effect of operating and design
Pr

variables on breakage rates using the mechanistic ball mill model, Miner.

Eng. 43 (2013) 91-101.


al

35. G.K.P. Barrios, R.M. de Carvalho, L.M. Tavares, Extending breakage


rn

characterisation to fine sizes by impact on particle beds, Trans. Instn. Min.


u

Metall. C, 120 (2011) 37-44.


Jo

36. F.N. de Magalhães, L.M. Tavares, Rapid ore breakage parameter estimation

from a laboratory crushing test, Int. J. Miner. Process. 126 (2014) 49-54.

37. M. Lindqvist, C.M. Evertsson, Improved flow- and pressure model for cone

crushers, Miner. Eng. 17 (2004) 1217-1225.

38. E. Hulthén, C.M. Evertsson, Real-time algorithm for cone crusher control with

two variables, Miner. Eng. 24 (2011) 987-994.

39. L.M. Tavares, R.D. Kallemback, Grindability of binary ore blends in ball mills,

Miner. Eng. 41 (2013) 115-120.

29
Journal Pre-proof

40. M.P. Bueno, T. Kojovic, M.S. Powell, F. Shi, Multi-component AG/SAG mill

model, Miner. Eng. 43-44 (2013) 12-21.

f
oo
pr
e-
Pr
al
u rn
Jo

30
Journal Pre-proof

f
oo
Figure 1. Laboratory-scale cone crusher used in experiments (left) and modeled version
pr
adopted in simulations (right)
e-
Pr
al
u rn
Jo

31
Journal Pre-proof

f
oo
pr
Figure 2. Cutaway view of the laboratory-scale cone crusher adopted in the simulations
e-
Pr
al
u rn
Jo

32
Journal Pre-proof

Closed side Open side


setting setting

f
oo
Figure 3. Illustration of a parallel view of the closed side setting of 5.0 mm, represented
by the internal circle in green (left) and the open side setting of 5.45 mm, represented by
the external circle in blue (right)
pr
e-
Pr
al
u rn
Jo

33
Journal Pre-proof

Figure 4. Snapshot of a simulation of the laboratory cone crusher (left), showing the
crusher discharge, illustrating flaky material leaving the crusher without undergoing
further rebreakage (right). Crusher with a CSS of 5 mm, a frequency of 616 rpm, a
throw of 0.45 mm and being fed with granulite particles sized from 16 mm to 22.4 mm.
Particles are colored according to their size

f
oo
pr
e-
Pr
al
u rn
Jo

34
Journal Pre-proof

2.5
Copper ore
Granulite
Limestone
Throughput (t/h) 2.0

1.5

1.0

0.5

f
0.0

oo
0 5 10 15 20
Simulation time (s)

5.0
pr Copper ore
e-
Granulite
4.0 Limestone
Pr
Power draw (kW)

3.0
al

2.0
rn

1.0
u

0.0
Jo

0 5 10 15 20
Simulation time (s)

Figure 5. Crushing simulations involving copper ore, granulite and limestone particles
at a frequency of 616 rpm, a CSS of 5 mm and a stroke of 0.45 mm. Discharge rates,
with throughput analysis starting at 2 seconds of simulation (top), and power draw
(bottom).

35
Journal Pre-proof

100 100
Cumulative passing (%)

Cumulative passing (%)


80 80

60 60

40 40

Copper ore Copper ore


20 Granulite 20 Granulite
Limestone Limestone
Feed Feed
0 0
0 5 10 15 20 25 0 5 10 15 20 25
Particle size (mm) Particle size (mm)

f
oo
Figure 6. Comparison between experimental (left) and simulated (right) particle size
distributions of crushing tests for the different materials studied. Crusher operating at a
frequency of 616 rpm, a CSS of 5 mm and a stroke of 0.45 mm.
pr
e-
Pr
al
u rn
Jo

36
Journal Pre-proof

4.0
Copper ore
Granulite
Limestone
Throughput (t/h) 3.0

2.0

1.0

0.0

f
0 2 4 6 8 10 12

oo
Closed side setting (mm)

3.0 pr Copper ore


Granulite
e-
2.5
Limestone
Specific energy (kWt/h)

2.0
Pr

1.5
al

1.0
rn

0.5
u

0.0
0 2 4 6 8 10 12
Jo

Closed side setting (mm)

Figure 7. Simulations on the effect of closed side setting on the crusher throughput (top)
and specific energy (bottom). Crusher operating at a frequency of 616 rpm and a stroke
of 0.45 mm.

37
Journal Pre-proof

1.0 4.0

0.8

Specific energy (kWh/t)


Throughput (t/h) 3.0

0.6
2.0
0.4

1.0
0.2
Throughput
Specific energy
0.0 0.0

f
0 500 1000 1500

oo
Crusher speed (rpm)

Figure 8. Effect of crusher speed on throughput and specific energy for copper ore
pr
simulations. Crusher operating at a CSS of 5 mm and a stroke of 0.45 mm.
e-
Pr
al
urn
Jo

38
Journal Pre-proof

3.0 3.0

2.5 2.5

Specific energy (kWh/t)


Throughput (t/h)
2.0 2.0

1.5 1.5

1.0 1.0

0.5 Throughput 0.5


Specific energy
0.0 0.0

f
0 0.2 0.4 0.6 0.8 1 1.2

oo
Crusher throw (mm)

Figure 9. Effect of size of throw on throughput and specific energy for copper ore
pr
crushing simulations. Crusher operating at a frequency of 616 rpm and a CSS of 5 mm.
e-
Pr
al
urn
Jo

39
Journal Pre-proof

100

Cumulative passing (%) 80

60

40

20 0.45 mm
0.7 mm
0.95 mm
0

f
0 2 4 6 8 10 12

oo
Particle size (mm)

Figure 10. Effect of size of stroke on the size distribution of crusher discharge for
pr
copper ore crushing simulations. Crusher operating at a frequency of 616 rpm and a
CSS of 5 mm.
e-
Pr
al
urn
Jo

40
Journal Pre-proof

3.0 3.0
Throughput
2.5 Specific energy 2.5

Specific energy (kWh/t)


Throughput (t/h)
2.0 2.0

1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0

f
0 20 40 60 80 100

oo
Percentage of copper ore

Figure 11. Simulations on the effect of percentage of copper ore in the feed blend on the
pr
throughput and specific energy of the crusher. Crusher operating at a frequency of 616
rpm, a CSS of 5 mm and a stroke of 0.45 mm.
e-
Pr
al
urn
Jo

41
Journal Pre-proof

100

80
Cumulative passing (%)

60
0% Limestone
40 25% Limestone
50% Limestone
75% Limestone
20 100% Limestone
Feed

f
0 4 8 12 16 20 24

oo
Particle size (mm)

Figure 12. Effect of the percentage of limestone (soft component) in the blend in the
pr
product size distribution. Crusher simulated at a frequency of 616 rpm, a CSS of 5 mm
and a stroke of 0.45 mm.
e-
Pr
al
u rn
Jo

42
Journal Pre-proof

Table 1. Physical characteristics of the samples


Measure Unit Copper ore Limestone Granulite
3
Specific gravity kg/m 2,930 2,710 2,790
Bond crushability work index kWh/t 13.5 5.0 15.2
Bond ball mill work index kWh/t 14.7 7.5 10.4
JKDWT A*b - 39.3 772 38.4

f
oo
pr
e-
Pr
al
u rn
Jo

43
Journal Pre-proof

Table 2. Mass/volume parameters of the cone geometry


Parameter Units Values
Volume cm³ 6222
Mass kg 48.53
Moment of inertia kg.m² 1.304

f
oo
pr
e-
Pr
al
u rn
Jo

44
Journal Pre-proof

Table 3. Contact parameters adopted in the simulations


Type of contact
Parameters
Particle/Particle Particle/Steel Steel/Steel
Static friction 0.8 0.5 0.3
Dynamic friction 0.8 0.5 0.3
Tangential stiffness ratio 1 1 1
Restitution coefficient 0.3 0.3 0.3

f
oo
pr
e-
Pr
al
u rn
Jo

45
Journal Pre-proof

Table 4. Parameters adopted to model the particle shapes


Parameter Particle #1 Particle #2 Particle #3 Particle #4
Vertical aspect ratio 0.8 0.5 0.7 1.0
Horizontal aspect ratio 1.0 0.8 1.2 1.0
Number of corners 25 20 25 20
Super-quadric degree 3.0 4.0 2.3 2.0

f
oo
pr
e-
Pr
al
u rn
Jo

46
Journal Pre-proof

Table 5. Summary of particle breakage parameters of the breakage model adopted in the
simulations
Parameters Copper ore Granulite Limestone
𝐸∞ (J/kg) 213.5 130.7 7.01
𝑑𝑜 (mm) 8.07 1.10 100
𝜑 1.22 1.99 0.80
𝜎 0.799 0.903 0.801
α1.2/β1.2 0.51/11.95 0.43/10.26 0.19/7.78
α1.5/β1.5 1.07/13.87 0.92/10.74 0.56/7.51
α2/β2 1.01/8.09 1.31/9.15 0.78/5.55
α4/β4 1.08/3.03 1.18/2.97 1.12/3.01
α25/β25 1.01/0.53 0.93/0.49 1.17/0.54

f
α50/β50 1.03/0.36 0.92/0.39 1.43/0.40

oo
α75/β75 1.03/0.30 0.90/0.31 1.92/0.42
𝛾 5.0 5.4 5.4
𝐴 (%)
𝑏′
67.7
0.029
pr 47.5
0.027
53.3
0.033
e-
Pr
al
u rn
Jo

47
Journal Pre-proof

Table 6. Comparison between simulated and experimental results for throughput, power
draw and specific energy of cone crusher tests
Material Coper ore Granulite Limestone

Throughput Simulation 0.77 1.31 2.05


(t/h) Experimental 0.69 0.95 1.87

Power draw Simulation 1.50 1.10 0.49


(kW) Experimental 2.71 2.29 0.65

Specific energy Simulation 1.95 0.84 0.24


(kWh/t) Experimental 3.93 2.41 0.35

f
oo
pr
e-
Pr
al
u rn
Jo

48
Journal Pre-proof

CREDIT AUTHOR STATEMENT

Flávio Pereira André: software; validation; investigation; data curation; writing –


original draft; writing – review & editing; visualization

Luís Marcelo Tavares: conceptualization; methodology; validation; resources; writing


– original draft; writing – review & editing; supervision; project administration; funding
acquisition

f
oo
pr
e-
Pr
al
u rn
Jo

49
Journal Pre-proof

Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

f
oo
pr
e-
Pr
al
u rn
Jo

50
Journal Pre-proof

 Simulation of a laboratory-scale cone crusher using DEM with polyhedral particles


 Good agreement between measured and simulated capacity for three materials
 Reasonable agreement between measured and simulated specific energy and size
 Studied effect of CSS, eccentric throw and speed on crusher performance
 Crushing binary mixtures simulated, showing impact of tough component on
throughput

f
oo
pr
e-
Pr
al
u rn
Jo

51
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12

You might also like