You are on page 1of 9

Powder Technology 395 (2022) 226–234

Contents lists available at ScienceDirect

Powder Technology

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

Impact energy of particles in ball mills based on DEM simulations


and data-driven approach
C.T. Jayasundara, H.P. Zhu ⁎
School of Engineering, Design and Built Environment, Western Sydney University, Locked bag 1797, Penrith, NSW 2751, Australia

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

Article history: The discrete element method has been used to simulate the particle flow in a ball mill under different operating
Received 19 June 2021 conditions. The model was validated by comparing the simulated results of the flow pattern and input power
Received in revised form 2 September 2021 with those measured from a same-scale laboratory mill. The impact energy of the particles under different oper-
Accepted 20 September 2021
ating conditions was analysed in detail. The results showed that the impact energy was affected by the operating
Available online 24 September 2021
conditions of the mill and can be linked to the grinding rate for a given material. The correlation between impact
Keywords:
energy and grinding rate follows first-order grinding kinetics. Mill performance decreases with increasing mill
Discrete element method size. Furthermore, a data-driven machine learning framework has been proposed to predict the impact energy
Ball mill for different operating conditions. It was found that the prediction for mills with diameters of 2000 and
Impact energy 3000 mm based on the training model developed by mills with diameter less than 254 mm could be achieved
Machine learning with an accuracy of 80% and a correlation coefficient of 0.9. Through the combination of DEM simulation and
Grinding rate data-driven approach, the computing time required in the determination of impact energy for large scale mills
can be dramatically reduced.
© 2021 Elsevier B.V. All rights reserved.

1. Introduction [6] proposed a scale-up procedure based on the population balance


models (PBMs), which considers particle breakage, transport and classi-
Ball mill is one of the most commonly used mills for the crushing and fication. This method allows product size distribution prediction and,
grinding of mineral ore. It is generally used to grind material down to when properly normalized with respect to energy, provides an
the particle size of 20 to 75 μm and can vary in size from a small batch energy-product size distribution relation. The PBM has been demon-
mill up to a mill with outputs of hundreds of tonnes per hour. The strated to provide a much superior prediction of grinding performance
final stages of comminution are generally performed in ball mills than the empirical Bond work index model [5,7]. However, due to the
which can be operated under dry or wet conditions. Any slight improve- lack of information on mechanical interactions between particles (i.e.
ment in grinding efficiency will be of immense economic benefit as micro-scale interactions), assumptions are often required in the appli-
grinding is a low efficient (< 20%, typically) and power intensive pro- cation of the empirical methods.
cess which may account for up to 40% of the direct operating cost of a With the advancement in computing resources, a numerical model
mineral processing plant [1]. Thus, it is crucial to improve the grinding based on the discrete element method (DEM) has been used to simulate
process efficiency by selecting the optimum operating conditions. the motion of particles in grinding mills. The DEM treats the particle
The operation process of comminution has been studied in the past flow as an assembly of particles and computes the movement of each
based on data gathered from various ores and semi-empirical models particle according to Newton's second law of motion. The interactions
[2]. The pioneer of the study was Bond and the equation proposed by between particles and between particles and the drum in a system are
him is known as the Bond's grinding equation [3,4]. As the Bond equa- determined based on the well-established contact mechanics. Since
tion did not consider breakage kinetics, all the effects in grinding were Mishra and Rajamani conducted a DEM study of grinding, this method
lumped into the work index. To overcome the limits of the energy has been used to study different milling processes, including ball mills
based work index, the population balance method (PBM) was intro- [7–11] and high speed stirred mills [12,13]. While the DEM simulations
duced later and became popular in the 1970s [5]. Herbst and Fuerstenau have been used for small systems to implement breakage of ground par-
ticles [14–16], direct simulation of particle breakage in a relatively large
system, especially, at an industrial scale, is still beyond the capability of
current computing resources. Alternatively, microdynamic properties
⁎ Corresponding author.
and their link with the grinding rate have been studied using the
E-mail address: h.zhu@westernsydney.edu.au (H.P. Zhu).

https://doi.org/10.1016/j.powtec.2021.09.063
0032-5910/© 2021 Elsevier B.V. All rights reserved.
C.T. Jayasundara and H.P. Zhu Powder Technology 395 (2022) 226–234

DEM. Through this, the DEM simulation results can be indirectly applied where vi, ωi and Ii are, respectively, the translational velocity, angular
for particle breakage processes. Recent studies showed that stirred mills velocity and moment of inertia of the particle, g is the gravitational
under a dry condition follow first-order grinding kinetics and the rate acceleration, and Ri was a vector running from the centre of the
constant can be linked with impact energy [17,18]. Similar correlations particle to the contact point with its magnitude equal to particle
were also reported for planetary ball mills [19,20]. This method provides radius Ri. Fijn, Fijs and Mrij represent, respectively, the normal contact
a way to predict the grinding rate of a given material under given oper- force, tangential contact force and rolling friction imposed on particle i
ating conditions provided that the number of particles is within the by particle j. Here, we use the simplified Hertz–Mindlin and Deresiewicz
computing capability. However, there is no report on the study of the model where contact forces are given by [25–27].
correlation between impact energy and grinding rate with varying  pffiffiffi pffiffiffiqffiffiffiffiffi 
2 3 
mill sizes, which is an important issue for simulations of grinding pro- F nij ¼ E Rξn 2 −γn E R ξn vij  nij nij ð3Þ
cess of large mills. 3
Although a tremendous growth in computational power has been
observed in the past, transient simulations of dynamic systems, using and
physics based numerical methods such as the DEM are prohibitively ex- h    i
3
F sij ¼ −sgn ðξs Þμ F nij 1− 1− min ξs , ξs, max =ξs, max Þ2 ð4Þ
pensive for large practical applications. A great deal of opportunity ex-
ists if we can efficiently learn the behaviour of the small system and  
then forecast it in different operating conditions. However, based on where E = Y= 1−σ e 2 , and Y and σ
e are, respectively, Young's modulus
the smaller system, a simple linear regression cannot be established and Poisson ratio; ξn is the overlap between particles i and j; nij is a unit
for larger systems with the presence of multiple independent variables vector running from the centre of particle j to the centre of particle i; R ¼
and the complexity of scale dependant variables. This can be overcome R=2 for mono-sized particle; and vij is the relative velocity of particles i
by the machine learning (ML) technique. ML has emerged as one of the and j at contact. The normal damping constant, γn, is the material
most promising technology in the past decade due to its capability to property directly linked to the coefficient of restitution e. ξs and ξs,max
provide valuable insights into complex systems. It enables learning are, respectively, the total and maximum tangential displacements
from many features (independent input variables) along with advanced of particles during contact. The rolling friction is given by −min {μr|
feature engineering, which leads to robust predictions of systems with Fnij|, μr|ωnij|}ωnij, where μr is the rolling friction coefficient, and ωnij is the
complex phenomena. It has been shown that ML-based predictions component of the vector of the relative angular velocity of particles i
are more accurate than linear regression techniques [21,22]. The prom- and j in their contact plane [28].
ise of machine learning has been explored in a variety of scientific disci- Simulations were carried out in the ball mills with the same geome-
plines including computationally transient models [23]. Although a try and operating conditions as used in the literature [29]. The mill di-
limited number of machine learning models have been used in grinding ameters are 127 mm, 254 mm and 508 mm where the mill length to
mills to predict various process parameters, none of the work has been diameter ratio is 1.15. Each mill consists of eight lifters. Three levels of
reported to investigate the effect of operating conditions and scale on steel balls of 12.7 mm, 19.05 mm, 25.4 mm and 31.75 mm diameters
mill performance using machine learning methods. were used as the grinding media. The filling levels M* was taken as
The objective of this study is to link the grinding rate to impact en- 30%, 40% and 50% of full mill and the mill speed N* was selected as
ergy for different mill sizes and overcome the limitations on the deter- 0.5, 0.6 and 0.7 of the critical speed. The critical speed is the speed at
mination of the impact energy associated with DEM simulations for which a mill drum rotates such that balls are stick to the drum, which
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
large scale grinding mills by applying machine learning models. The ef- is given by 2g=ðD−dÞ where D and d are the mill diameter and parti-
fect of different operating conditions of the ball mill on particle flow was cle diameter in meters respectively. For this work, only the motion of
first investigated in terms of microdynamic variables related to flow grinding media was simulated without considering the ore charge.
structure such as velocity, impact energy and collision frequency. The The ore charge was considered entrapped in between colliding balls
relation between the grinding rate constant and impact energy was which were subjected to the collision forces. The parameters used in
then established. After that, exploratory data analysis was conducted the simulations are given in Table 1. Sliding friction and restitution coef-
to explore the correlations among the specific impact energy and oper- ficients affect the input power of particle systems [30]. To determine
ating conditions. Finally, the impact energy of large systems based on their values for the mill considered, different values have been
the predictive data model was studied. attempted until a comparative input power with the experiment [29]
was achieved. Particle density is the same as that in the experiment.
2. DEM model and simulation conditions Young's modulus was selected based on the previous simulations for
the soft sphere DEM model [31].
The DEM model used in this work is based on the soft-sphere model
which has been extensively used to study various phenomena, such as
Table 1
particle packing, transport properties, hopper flow, mixing and granula-
Physical parameters used in the simulations.
tion [24]. The soft-sphere model is more suitable for systems such as
grinding mills because the model is capable of handling more than Parameter Value
one contact at a time. The motion of particles is described by Newton's Ball density, ρ (kgm−3) 8050
laws of motion. The governing equations for the translational and rota- Young's modulus, Y (Nm−2) 2.0 × 107
tional motion of particle i with mass mi and moment of inertia Ii can be Poisson ratio, σ e 0.3
sliding friction coefficient, μs 0.4
written as:
Rolling friction coefficient, μr (d) 0.01

dvi   Restitution coefficient, e 0.68


mi ¼ ∑ F nij þ F sij þ mi g ð1Þ Particle diameter, d (mm) and maximum 12.7 (110000), 19.05 (120000),
dt number of particles used in the simulation 25.4 (69638), 31.75 (35000)
(given in the brackets)
and Mill diameter, D (mm) 127, 254, 508, 2000, 3000
Mill speed as a fraction of critical speed N* 0.5, 0.6, 0.7
dωi  
Fill level fraction M* 0.3, 0.4, 0.5
Ii ¼ ∑ Ri  F sij þ M rij ð2Þ
dt Numerical time step (s) 1.0 × 10−5

227
C.T. Jayasundara and H.P. Zhu Powder Technology 395 (2022) 226–234

In each simulation, particles were randomly generated in the milling learning model to identify the optimal parameter settings. The best
vessel and allowed to settle under gravity. The mill was then rotated method to determine the optimal settings is to try many different com-
clockwise and allowed to reach the steady state before data collection binations and evaluate the performance of each model. Under-fitting
began. In this work, the steady state was believed to be reached when occurs when the model does not fit the data well and is unable to cap-
the variation of space average particle velocity was insignificant. ture the underlying trend in which a low accuracy is observed in the
When the system was in a steady state, data collection started for training and test dataset. To the contrary, over-fitting occurs when the
two complete revolutions of the mill. Both particle-particle and model fits the data too well, capturing all the noises. In this situation,
particle-wall collisions were collected to calculate total impact energy high accuracy can be observed in the training dataset, whereas the
and collision frequency. The impact energy for a collision was defined same model will result in a low accuracy on the unseen dataset. There-
as the kinetic energy, given by ½mijv2ij, where mij = 2mimj/(mi + mj) fore, cross-validation has been used for the hyperparameter optimiza-
and vij(=|vi − vj|) is the relative normal collision velocity between tion here. This is a standard practice for hyperparameter optimization,
two contacted particles i and j. For the collision between a particle and whose detail can be found in [32].
the drum, the impact energy was calculated by ½mi|vi|2. A contact was
counted when the overlap occurred for the first time. Existing overlap
was not considered as a new contact until it separated and come into
3. Results and discussion
contact again.
The results obtained from the DEM simulations under different op-
3.1. Flow pattern
erating conditions have been used to develop the ML model. The mill
was characterized by 4 input features: mill diameter, mill rotating
Fig. 1 shows the particle flow of different size of mills when M* = 0.5
speed (percentage of critical speed), fill volume and particle size. The ef-
and N* = 0.6 at the steady state. The colour in the figure represents the
fect of material properties such as particle density, rolling friction coef-
particle velocity. Note that the DEM simulations are only for the grind-
ficient, sliding friction coefficient was not considered. As the input data
ing media without considering the ground (fine) particles. Compared
have different varying ranges, they were first normalized, before being
to the grinding media, the number of ground particles is significantly
fed into the model, by
less and the effect of the fine particles on the grinding media is insignif-
ðx−xmin Þ icant [20,33]. The motions of balls for different mill diameters are similar
xscaled ¼ ð5Þ at the same fraction of critical speed, which is identical to that observed
ðxmax −xmin Þ
by other scholars [33,34,37] for a tumbling ball mill. The balls near the
where x, xmin and xmax are the original feature and the minimum and bottom are lifted by the lifters and dragged up around the liner. They
maximum values of the original feature, respectively. lose contact with the drum in the second quadrant and most of the
The accuracy of the model is determined by the mean absolute per- balls then flow down the steep free surface into the horizontal toe re-
centage error gion. This is generally described as the cascading part of the flow.

 Once they reach the lower region of the drum, the balls again move
1 n y −z up towards the upper region. The particles reach a peak velocity just
accuracy ¼ 100∗ 1− ∑ i i ð6Þ
n i¼1 yi below the centre of the mill. The next highest velocity cluster of particles
are the ones rotating with the mill near the shell. There is a semi-circular
where yi, zi and n is the target value obtained from the DEM, predicted band indicated by the light blue of particles that move quite slowly. This
value from the model and the sample size, respectively. The correlation behaviour is generally observed in ball mills when they operate below
coefficient between two variables ai and bi is given by the critical speed [34]. When the mill size increases, the particles are
  raised to a higher elevation and then drop under gravity, promoting a
∑ðai −aÞ bi −b higher velocity.
r ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2 ð7Þ
To further examine the effect of mill size on particle velocity, Fig. 2
∑ðai −aÞ2 ∑ bi −b shows the probability distribution (PDF) of particle velocity at different
critical speed fractions. The profiles of the distribution are similar for all
where a and b are the means of ai and bi, respectively. cases considered. It increases, reaches a peak, and then decreases with
Four different popular ML methods, including linear regression (LR), the increase of the velocity. When mill size increases distribution be-
partial least square regression (PLS), elastic net regression (EN) and ran- comes wider suggesting more particles with higher velocity. When
dom forest regression (RF), were tested to build the predictive model. the mill size increases, the particle velocity increases because of a high
Hyperparameter tuning has been used to improve the accuracy of the percentage of particles having higher velocity in the middle region
models. Hyperparameter tuning is an essential aspect of the machine close to the centre of the mill where the cascading flow can be observed.

Fig. 1. Flow patterns in ball mills with different sizes: (a) 127 mm; (b) 254 mm; (c) 508 mm; (d) 2000 mm; and (e) 3000 mm, when d = 25.4 mm, M* = 0.5 and N* = 0.6.

228
C.T. Jayasundara and H.P. Zhu Powder Technology 395 (2022) 226–234

Fig. 2. Probability distribution and mean velocity for different mill sizes at different fractions of critical speed: (a) N* = 0.5; (b) N* = 0.6; and (c) N* = 0.7. v is the mean velocity of particles.

When mill speed increases particles obtain higher velocity. As a result, increases at a higher rate than total impact energy in the system.
mean velocity increases with increasing mill size or mill speed. Fig. 4b shows the effect of critical speed and fill volume on Ei. for given
particle size and mill size. When filling volume increases, Ei decreases.
This is because the total mass increase is higher than the rate of
3.2. Impact energy, collision frequency and input power

When the mill is in operation, impact energy between particle-


particle and particle-wall contributes to the liberation of fine powder.
The amount of release of the fine powder depends on the intensity
and the frequency of the impact energy. Fig. 3a shows the variation of
collision frequency with impact energy per unit time for different
sizes of mills when d = 25.4 mm, M* = 0.5 and N* = 0.6. The impact
energy per unit time was calculated by dividing the total impact energy
by the simulation time. Both particle-particle and particle-wall colli-
sions were divided into bins of different values of impact energy, and
each energy class was plotted as collisions per second which give the
collision frequency. Generally, particle-particle impact energy is smaller
in magnitude than particle-wall impact energy because the bulk of the
particles are closely packed which limits the relative velocity among
particles. Particle-wall impacts tend to have higher impact energies
caused by the cataracting particles which fall onto the toe region with
a higher velocity. A high number of particle-particle collisions contrib-
ute to the smooth variation whereas a low number of particle-wall col-
lisions contribute towards the tail where a spiky variation can be
observed. When milling size increases, distribution shifts to right sug-
gesting that total impact energy increases with increasing mill size be-
cause of an increased number of particles in larger mills. However,
impact energy per particle and collision frequency per particles de-
crease with increasing mill size (Fig. 3b). Although a high number of im-
pacts are observed at low energies, all the impacts may not be useful for
feed powder breakage. Impact energy should be high enough to initiate
crack propagation even under repeated impacts [18]. For each ground
material, there exists a threshold impact energy which is a function of
particle size and material properties. The work reported by Capece
et al. [18] observed that the size-specific impact energy threshold was
calculated to be 2.64 × 10−5 J based on the size-independent energy
threshold. In this analysis, we did not consider impacts less than
2.5 × 10−5 J as useful impacts.
The impact energy is affected by the operating conditions of the mill.
To illustrate this, Fig. 4a shows the effect of particle size and mill size on
specific impact energy Ei for a given critical speed (N* = 0.6) and fill
volume (M* = 0.5). The specific impact energy of particles is the total
impact energy per unit time divided by the particle mass. Note that
two additional cases have been included when considering the effect
of each property except particle diameter to show the trend of
variation more clearly. When particle size increases, Ei increases
because of increased individual particle mass which contributes to the Fig. 3. Impact energy and collision frequency for different size of mills when d = 25.4 mm,
calculation of impact energy among particles. When mill size M* = 0.5 and N* = 0.6: (a) impact energy per unit time and collision frequency; and
increases, Ei decreases because of total particle mass in the system (b) variations of impact energy and collision frequency per particle with mill size.

229
C.T. Jayasundara and H.P. Zhu Powder Technology 395 (2022) 226–234

product of the total torque with angular mill speed (in rad/s), and it
has an impulsive nature. Here, the average power was calculated and
compared with the experimental power draw.
Fig. 5a shows the comparison of specific input power Pi (i.e., the
input power per unit time per unit mass) based on the simulations ob-
tained from the experiments by Malghan [29]. The values of Pi from
simulations and experiments for different sized mills are very close.
There is a linear relationship between specific input power and mill
size in log-log scale, which indicates that Pi is proportional to Dn
where n = 2.6 which is within the range reported in the experiments
[3,35]. The result suggests that input power obtained from small scale
mills can be extrapolated for large scale mills. The good agreement
with the experimental values, combined with the qualitative
agreements with previous studies [33,34,37] in terms of flow patterns
discussed in Section 3.1, demonstrates the validity of the DEM model.
A more extensive validation of the DEM code used in this work has
been conducted in our previous study [31]. Fig. 5b shows the ratio of
specific impact energy to specific input power γ for different mill
sizes. γ gives a measure of grinding performance [17]. When mill size in-
creases, γ decreases, which indicates that the rate of increase of Pi is
higher than the rate of increase of Ei. This is because the bulk of the
material need to be lifted to a higher elevation for a larger mill size, so
the energy per unit mass required is higher. As a result, when mill size
is increased, Pi increases at a higher rate than Ei. The results suggest
that under the same operating conditions (when M* and N* are
identical), grinding performance decreases with increasing mill size.
Other operating conditions with different fill volumes and critical
speeds also follow the same trend. However, larger mills have higher
throughput than smaller mills because of increased mill capacity.
Therefore, the selection of an optimum operating condition is a
compromise between mill performance and throughput.

3.3. Grinding rate constant and impact energy relation

Experiments reported by Malghan [29] for different sizes of ball mills


were analysed to obtain the grinding rate of calcium carbonate as
ground material and steel balls as grinding media. Cumulative particle
size distribution curve was plotted and used to obtain the particle size
for 50% cumulative passing level for a given time which was known as
D50,t. This procedure was repeated for all the experiments. When
grinding time t increases, the rate of size reduction decreases. The
relationship between size reduction and grinding time can be
approximated by an exponential function. When grinding time
increases, Dt/Dt0 tends to level off reaching the lowest value which is
known as the grinding limit Dt∞. Here Dt is the particle size at time t
and Dt0 is the initial particle size. Then the relationship between
(D50,t − D50,∞)/(D50,0 − D50,∞) vs grinding time t can be given as
ðD50,t −D50,∞ Þ
= e−Kpt, where Kp is known as grinding rate [17,36].
ðD50,0 −D50,∞ Þ
Fig. 4. Variation of specific impact energy: (a) with mill size when M* = 0.5, N* = 0.6, and
d = 25.4 mm, and with particle size when M* = 0.5, N* = 0.6 and D = 254 mm; (b) with Fig. 6 shows log-linear plot of (D50,t − D50,∞)/(D50,0 − D50,∞) vs
percentage of critical speed when d = 25.4 mm, D = 254 mm and M* = 0.5, and with fill grinding time t which shows straight lines. Though different slopes are
volume when d = 12.4 mm, D = 254 mm and N* = 0.6. observed for different mill sizes under different operating conditions,
the grinding rate increases with increasing mill size. This confirms
that the grinding process of the ball mill follows the first-order kinetics,
increase in the total impact energy. When increasing critical speed, and the particle size decays exponentially with time. Similar observa-
relative velocity among particles increases. As a result, higher critical tions were reported for dry and wet grinding of different mills such as
speed promotes higher Ei. the stirred mill and planetary ball mill [17,20].
The energy required to drive the mill varies depending on the ore While DEM modelling of particle flow provides a significant amount
type, properties of grinding media and operating conditions such as of information at the particle scale, simulating particle breakage for an
mill speed and fill volume, which means that input power is system de- actual grinding process under industrial-scale large systems consisting
pendant. Mill input power was calculated based on the torque gener- of trillions of fine particles is beyond the capability of current computing
ated by those particles which were colliding with the mill drum. For resources. Alternatively, establishing a correlation between impact en-
each time step, individual torques were summed from the contributing ergy and grinding rate provides a way of eliminating the simulation of
particles to obtain the instantaneous torque on the mill. Those particles fine particles. In the previous studies [17,19,20,36], it has been shown
which are not in contact (free flight) do not contribute to torque until that there is a link between grinding rate and impact energy for given
they meet the mill drum. Instantaneous power consumption is the grinding materials under different mill speeds and fill volume. However,

230
C.T. Jayasundara and H.P. Zhu Powder Technology 395 (2022) 226–234

Fig. 5. (a) Comparison between the simulations and experiments [29] in terms of specific input power vs mill size; and (b) the variation of the ratio of specific impact energy to input power
with mill size for different N* when M* = 0.5 and d = 25.4 mm.

this correlation has not been examined for ball mills with varying mill
sizes and other operating conditions. Fig. 7 shows the correlation be-
tween specific impact energy and grinding rate in a log-log plot based
on the simulations of the three sized mills. Specific impact energy has
been noted by researchers in comminution as a major factor in rock
fracture. High specific impact energy promotes a higher grinding rate
[37]. The correlation suggests that grinding rate can be predicted for
any given mill size provided that the specific impact energy is known.
However, as mentioned above, simulating a large number of particles
using physics-based DEM modelling remains prohibitively expensive.
Alternatively, a data-driven approach can be utilized to determine im-
pact energy for large systems.

3.4. Correlations among input features and impact energy

Exploratory data analysis is an important step in the machine learn-


ing model because it can help identify obvious errors, better understand
patterns within the data, detect outliers or anomalous events, and find
Fig. 7. Correlation between specific impact energy and grinding rate for different operating
interesting relations among the variables. This procedure mainly exam-
conditions: mill size, mill speed, fill volume and particle size.
ines the correlation coefficient among input features and the

distribution of each input feature. Table 2 shows the correlations


among input features and specific impact energy Ei. Generally,
correlation coefficient r measures the degree of linearity between two
variables and it varies between −1 to 1. A positive correlation means
a direct linear relationship and a negative correlation means an inverse
linear relationship between the two variables. Positive correlations be-
tween N* and Ei, and d and Ei suggest that increase in mill speed or
particle diameter would increase the specific impact energy. Negative
correlations between M* and Ei, and D and Ei suggest that increase in
fill volume or mill diameter would decrease the specific impact

Table 2
Correlation among input features and specific impact energy.

d D M* N* Ei

d 1 0.071 0.047 −0.005 0.307


D 0.071 1 0.198 −0.058 −0.247
M* 0.047 0.198 1 −0.013 −0.630
N* −0.005 −0.058 −0.013 1 0.609
Fig. 6. Normalized particle sizes of calcium carbonate ground samples as a function of
Ei 0.307 −0.247 −0.630 0.609 1
grinding time for different mill sizes when M* = 0.5 and N* = 0.6.

231
C.T. Jayasundara and H.P. Zhu Powder Technology 395 (2022) 226–234

energy. Since all the four input features show low r (<0.7) with Ei, it 3.5. Impact energy of large systems based on the predictive data model
can be expected that these input features have nonlinear relationships
with Ei. Due to the computing cost associated with the acquisition of DEM
Furthermore, the effect of input features on specific impact energy data, we investigated the minimum size of training data required for
has been examined through box plots given in Fig. 8. Box plots show reasonable accuracy. Dataset based on 108 cases was randomly divided
the distribution of the specific impact energy with each variable. The into different subsets of training data varying from 43, 54, 64, 76, 86 and
box ranges from the first quartile (Q1) to the third quartile (Q3) of the 97 which approximately correspond to 40%, 50%, 60%, 70%, 80% and 90%
distribution, and the range represents the interquartile range (IQR). of the total data, respectively. The trained models were used to predict
The median is indicated by a line across the box. The whiskers on the specific impact energy for large mills of D = 2000 mm and
the box plot extend from Q1 and Q3 to the most extreme data points 3000 mm with three levels of d, N* and M* of each operating condition
which give the range of the data points. If the median is in the middle which gives 54 different cases. Then the predicted values for D ≥ 2000
of a box, and the upper and lower whiskers have the same length, the were compared with the simulation ones, and the accuracy for each
distribution is symmetric. It can be seen that most of the box plots case was calculated by using Eq. 6. Fig. 9 shows the accuracy of different
show asymmetric distributions, which suggests that data may have number of training data for the four models. None of the models showed
complex characteristics. An increase and then decrease in median a significant improvement in the accuracy beyond 86 training data,
values of d and D suggest a strong nonlinearity with the specific im- which indicates that 86 data for training and 22 data for testing can be
pact energy. However, the median value of M* shows a decrease used for the model development. Among the four methods, random for-
whereas that of N* shows an increase, which suggests that M* and est performs the best, with an accuracy of 80% for the prediction of spe-
N* have a weaker nonlinear correlation with E i, than d and D. cific power for large scale mills (D ≥ 2000 mm).
Therefore, the values of r of M* and N* are higher than those of d Fig. 10 shows the scatter plot of comparison between actual and pre-
and D. The exploratory data analysis based on Table 2 and Figs. 8 dicted specific impact energy based on the four ML methods when 86
showed that the correlations between the specific impact energy data sample was used for training. When D ≤ 508 mm, all four models
and input features were nonlinear. The selected machine learning show data points close to the dotted line. However, when D = 2000
model should be able to capture the nonlinear behaviour with or 3000 mm, models LR, PLS and EN show scattered data which results
acceptable accuracy. in a low r (= 0.27 for LR and PLS, and 0.28 for EN). Compared to the

Fig. 8. Distribution of specific impact energy with each input feature: (a) particle diameter; (b) mill diameter; (c) mill speed; (d) fill volume.

232
C.T. Jayasundara and H.P. Zhu Powder Technology 395 (2022) 226–234

three models, RF shows a significantly higher value of r (= 0.9). The re-


sults given in Figs. 9 and 10 suggest that random forest model gives a
reasonable prediction up to sixfold mill sizes while LR, PL and EN models
are not suitable for predicting extrapolated values. Random forest com-
bines several decision trees and is based on randomized feature selec-
tion [38], which causes that it is capable of handling complex systems
over other models. Moreover, nonlinear nature of random forest can
give it an advantage over other algorithms for the considered system
with nonlinear correlations between specific impact energy and input
features as discussed in Section 3.4. These features may be the reasons
why random forest is much more accurate than other models. It is
noted that further increase of mill size may not perform well, which is
a limitation of the proposed model.
Table 3 shows DEM simulation time for different mill sizes when the
system is at a steady state. Computational time significantly increases
with the number of particles. For instance, the computational time for
D = 508 mm takes about 6 min whereas D = 3000 mm takes about
800 min for a two-second simulation. Therefore, prediction based on
Fig. 9. Comparison of accuracy of different number of training data for linear regression
the RF model is a tremendous improvement because it takes only
(LR), partial least square regression (PLS), elastic net regression (EN), and random forest milliseconds to obtain the results. The other advantage of this model is
regression (RF). that it can be used to predict impact energy for any different operating

Fig. 10. Machine learning predictions of specific impact energy based on different methods: (a) linear regression; (b) partial least square regression; (c) elastic net regression; and
(d) random forest regression. The red dots represent the test data whereas blue dots represent a larger mill with D = 2000 or 3000 mm.

233
C.T. Jayasundara and H.P. Zhu Powder Technology 395 (2022) 226–234

Table 3 [3] F.C. Bond, The 3rd theory of comminution, Trans. Am. Inst. Min. Metallurg. Eng. 193
Comparison of computational time for different mill sizes when d = 25.4 mm and (1952) 484–494.
M* = 0.5. [4] F.C. Bond, Confirmation of the 3rd theory, Trans. Am. Inst. Min. Metallurg. Eng. 217
(1960) 139–153.
D (mm) Np DEM simulation time (minutes) [5] J.A. Herbst, G.A. Grandy, T.S. Mika, On the development and use of lumped parame-
ter models for continuous open- and closed-circuit grinding systems, Miner. Pro-
254 400 2 cess. Extr. Metall. 80 (1971) C193–C198.
508 1874 6 [6] J.A. Herbst, D.W. Fuerstenau, Scale-up procedure for continuous grinding mill design
2000 30,300 300 using population balance models, Int. J. Miner. Process. 7 (1980) 1–31.
3000 69,630 800 [7] J.A. Herbst, A microscale look at tumbling mill scale-up using high fidelity simula-
tion, Int. J. Miner. Process. 74 (2004) S299–S306.
[8] B.K. Mishra, R.K. Rajamani, The discrete element method for the simulation of ball
mills, Appl. Math. Model. 16 (1992) 598–604.
[9] P.W. Cleary, Predicting charge motion, power draw, segregation and wear in ball
combinations provided that the model descriptors are well within the
mills using discrete element methods, Miner. Eng. 11 (1998) 1061–1080.
training data set. Prediction based on 508 mm up to 3000 mm is a signif- [10] P.W. Cleary, Charge behaviour and power consumption in ball mills: sensitivity to
icant achievement because simulation time for 3000 mm is 133 times mill operating conditions, liner geometry and charge composition, Int. J. Miner. Pro-
the simulation time of D = 508 mm. The framework proposed in this cess. 63 (2001) 79–114.
[11] M.S. Powell, N.S. Weerasekara, S. Cole, R.D. LaRoche, J. Favier, DEM modelling of liner
work will significantly reduce the computational time when predicting evolution and its influence on grinding rate in ball mills, Miner. Eng. 24 (2011)
impact energy for large systems and will be a promising technique for 341–351.
large scale DEM simulations. [12] C.T. Jayasundara, R.Y. Yang, A.B. Yu, D. Curry, Discrete particle simulation of particle
flow in IsaMill—effect of grinding medium properties, Chem. Eng. J. 135 (2008)
103–112.
4. Conclusions [13] R.Y. Yang, C.T. Jayasundara, A.B. Yu, D. Curry, DEM simulation of the flow of grinding
medium in IsaMill, Miner. Eng. 19 (2006) 984–994.
[14] P.W. Cleary, C.Z.C. Raymond, M. Prakash, M.D. Sinnott, S.M. Harrison, S. Mead, Pre-
Numerical modelling based on DEM has been performed on ball diction of industrial, biophysical and extreme geophysical flows using particle
mills of different sizes. It has been shown that the grinding of ball methods, Eng. Comput. 30 (2013) 157–196.
mills of different sizes follows the first-order kinetics. Although different [15] G.W. Delaney, P.W. Cleary, M.D. Sinnott, R.D. Morrison, Novel application of DEM to
modelling comminution processes, 9th World Congress on Computational
collision environments exist with different sizes of mills, impact energy MechanicsSydney, 2010 , Australia.
is a useful index that determines the grinding rate. Regardless of the mill [16] D.O. Potyondy, Cundall, A bonded-particle model for rock, Int. J. Rock Mech. Min. Sci.
size, for a given material, a unique correlation exists between impact en- 41 (2004) 1329–1364.
[17] C.T. Jayasundara, R.Y. Yang, A.B. Yu, J. Rubenstein, Effects of disc rotation speed and
ergy and the grinding rate constant in which the grinding rate increases
media loading on particle flow and grinding performance in a horizontal stirred
with increasing mill size. When increasing the mill size under similar mill, Int. J. Miner. Process. 96 (2010) 27–35.
dynamic conditions, the rate of increase of specific input energy is [18] M. Capece, E. Bilgili, R. Dave, Insight into first-order breakage kinetics using a
higher than the rate of increase in specific impact energy. As a result, particle-scale breakage rate constant, Chem. Eng. Sci. 117 (2014) 318–330.
[19] J. Kano, H. Mio, F. Saito, Correlation of grinding rate of gibsite with impact energy of
grinding performance decreases with increasing mill size. However, in ball mills, Am. Inst. Chem. Eng. J. 46 (2000) 1694–1697.
terms of throughput large mills are always the best choice because of in- [20] J. Kano, F. Saito, Correlation of powder characteristics of talc during planetary ball
creased mill capacity. The impact energy is affected by the operating milling with the impact energy of the balls simulated by the particle element
method, Powder Technol. 98 (1998) 166–170.
conditions of the mill. The specific impact energy increases with an in- [21] D.T. Ahneman, J.G. Estrada, S. Lin, S.D. Dreher, A.G. Doyle, Predicting reaction perfor-
crease in particle size or percentage of critical speed, or with a decrease mance in C-N cross-coupling using machine learning, Science 360 (2018) 186–190.
in mill size or fill volume. The exploratory data analysis showed that the [22] V.A. Huynh-Thu, A. Irrthum, L. Wehenkel, P. Geurts, Inferring regulatory networks
from expression data using tree-based methods, PLoS One 5 (2010), e12776.
correlations between impact energy and input features were nonlinear. [23] P. Kumar, K. Sinha, K.N. Nandkishor, Y. Shin, H. Raimundo, B.M. Laurie, Y.S. Ahmad, A
DEM combined with machine learning showed a novel technique for machine learning framework for computationally expensive transient models, Nat.
predicting the grinding rate of large-scale mills based on the data gener- Res. Forum 10 (10) (2020) 11492.
[24] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of particulate sys-
ated from small scale numerical models. Among different machine tems: theoretical developments, Chem. Eng. Sci. 62 (2007) 3378–3392.
learning methods, random forest performed the best with r = 0.9 and [25] N.V. Brilliantov, F. Spahn, J.M. Hertzsh, T. Poshel, Model for collisions in granular
an accuracy of 80% for the prediction of specific power for large scale gases, Phys. Rev. E 53 (1996) 5382.
[26] K.L. Johnson, Contact Mechanics, Cambridge University Press, 1985.
mills (D ≥ 2000 mm). This is a significant improvement in simulating
[27] C.T. Jayasundara, R.Y. Yang, B.Y. Guo, A.B. Yu, J. Rubenstein, Effect of slurry properties
large-scale mills where numerical modelling has limitations due to a on particle motion in IsaMills, Miner. Eng. 22 (2009) 886–892.
large number of particles. The results show that relatively high accuracy [28] H.P. Zhu, A.B. Yu, The effects of wall and rolling resistance on the couple stress of
can be expected up to sixfold mill sizes. The framework described in this granular materials in vertical flow, Phys. A - Stat. Mech. Appl. 325 (2003) 347–360.
[29] S.G. Malghan, The Scale-Up of Ball Mills Using Population Balance Models, Univer-
work in principle can be applied for other types of mills as well. sity of California, Berkeley, 1976.
[30] W.J. Rong, Y.Q. Feng, P. Schwarz, T. Yurata, P. Witt, B.K. Li, T. Song, J.W. Zhou, Sensi-
Declaration of Competing Interest tivity analysis of particle contact parameters for DEM simulation in a rotating drum
using response surface methodology, Powder Technol. 362 (2020) 604–614.
[31] C.T. Jayasundara, R.Y. Yang, A.B. Yu, D. Curry, Discrete particle simulation of particle
The authors declare that they have no known competing financial flow in IsaMill, Ind. Eng. Chem. Res. 19 (2006) 6349–6359.
interests or personal relationships that could have appeared to influ- [32] R.W. Kennard, L.A. Stone, Computer aided design of experiments, Technometrics 11
(1969) 1537–2723.
ence the work reported in this paper. [33] T. Iwasaki, T. Yabuuchi, H. Nakagawa, S. Watano, Scale-up methodology for tum-
bling ball mill based in impact energy of grinding balls using discrete element anal-
Acknowledgements ysis, Adv. Powder Technol. 21 (2010) 623–629.
[34] P.W. Cleary, M. Prakash, M.D. Sinnott, M. Rudman, R. Das, in: E. Onate, R. Owen
(Eds.),Large scale simulation of industrial, engineering and geophysical flows
The authors are grateful to the ARC Research Hub for Computational using particle methods, Particle-Based Methods: Fundamentals and Applications,
Particle Technology (IH140100035) and Western Sydney University for Springer, Dordrecht Netherlands, 2011, pp. 89–111.
the financial support of this work. [35] C.A. Rowland, Comparison of work indices calculated from operating data with
those from laboratory test data, Proceedings Tenth International Minerals Process-
ing Congress, The Institution of Mining and Metallurgy, London 1973, pp. 47–61.
References [36] C.T. Jayasundara, R.Y. Yang, A.B. Yu, Effect of the size of media on grinding perfor-
mance in stirred mills, Miner. Eng. 33 (2012) 66–71.
[1] B.A. Wills, Mineral Processing Technology, 5th ed., 1992 Oxford. [37] N.S. Weerasekara, L.X. Liu, M.S. Powell, Estimating energy in grinding using DEM
[2] N.S. Weerasekara, M.S. Powell, P.W. Cleary, L.M. Tavares, M. Evertsson, R.D. modelling, Miner. Eng. 85 (2016) 23–33.
Morrison, J. Quist, R.M. Carvalho, The contribution of DEM to the science of commi- [38] L. Breiman, Random forests, Mach. Learn. 45 (2001) 5–32.
nution, Powder Technol. 248 (2013) 3–24.

234

You might also like