You are on page 1of 89

DEM TECHNICAL MANUAL

dem technical manual i


ROCKY

Copyright

copyright ©2020, esss. all rights reserved.

No part of this documentation may be reproduced in any form, by any means,


without the prior written permission of ESSS.

ESSS makes no representations or warranties with respect to the program material


and specifically disclaim any implied warranties, accuracy, merchantability or fitness
for any particular purpose. Furthermore, ESSS reserves the right to revise the
program material and to make changes therein without obligation to notify purchaser
of any revisions or changes except specific errors determined to be incorporated in the
program material. It shall be the responsibility of ESSS to correct any such errors in
an expeditious manner. In no event shall ESSS be liable for any incidental, indirect,
special, or consequential damages arising out of the purchaser’s use of program
material.

© 2020, esss - all rights reserved


dem technical manual ii
ROCKY

Contents

1 Introduction to DEM 1

2 Physical models in Rocky 3

2.1 Contact force models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3


2.1.1 Normal force models . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.1.1 Hysteretic linear spring model . . . . . . . . . . . . 4
2.1.1.2 Linear spring-dashpot model . . . . . . . . . . . . . 7
2.1.1.3 Hertzian spring-dashpot model . . . . . . . . . . . 9
2.1.2 Tangential force models . . . . . . . . . . . . . . . . . . . . . . 10
2.1.2.1 Linear spring Coulomb limit model . . . . . . . . . 10
2.1.2.2 Coulomb limit model . . . . . . . . . . . . . . . . . 11
2.1.2.3 Mindlin-Deresiewicz model . . . . . . . . . . . . . . 12

2.2 Adhesive force models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13


2.2.1 Constant adhesive force model . . . . . . . . . . . . . . . . . . 13
2.2.2 Linear adhesive force model . . . . . . . . . . . . . . . . . . . . 14
2.2.3 Leeds adhesive force model . . . . . . . . . . . . . . . . . . . . 16
2.2.4 JKR adhesive force model . . . . . . . . . . . . . . . . . . . . . 18

2.3 Rolling resistance models . . . . . . . . . . . . . . . . . . . . . . . . . . . 19


2.3.1 Rolling resistance type 1 . . . . . . . . . . . . . . . . . . . . . . 19
2.3.2 Rolling resistance type 3 . . . . . . . . . . . . . . . . . . . . . . 19

2.4 Breakage models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21


2.4.1 Ab-T10 breakage model . . . . . . . . . . . . . . . . . . . . . . 21
2.4.2 Tavares breakage model . . . . . . . . . . . . . . . . . . . . . . 23
2.4.3 Breakage for large deformations . . . . . . . . . . . . . . . . . 26
2.4.4 Fragment size distribution models . . . . . . . . . . . . . . . . 27
2.4.4.1 Gaudin-Schumann . . . . . . . . . . . . . . . . . . . 28
2.4.4.2 Incomplete beta function . . . . . . . . . . . . . . . 28

2.5 Wear model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2.6 Thermal model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31


2.6.1 Contact heat transfer rate . . . . . . . . . . . . . . . . . . . . . 32

© 2020, esss - all rights reserved


dem technical manual iii
ROCKY

2.6.2 Joint heat transfer rate . . . . . . . . . . . . . . . . . . . . . . . 33


2.6.3 Thermal conduction correction models . . . . . . . . . . . . . . 35

2.7 Coarse-Grain Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36


2.7.1 Contact frictional forces . . . . . . . . . . . . . . . . . . . . . . 38
2.7.2 Adhesive forces . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.7.3 Contact heat transfer . . . . . . . . . . . . . . . . . . . . . . . . 39
2.7.4 Radl et al. kinetic energy dissipation model . . . . . . . . . . . 41

3 Particle types in Rocky 44

3.1 Fiber particle shapes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44


3.1.1 Flexible straight fibers . . . . . . . . . . . . . . . . . . . . . . . 44
3.1.2 Flexible custom fibers . . . . . . . . . . . . . . . . . . . . . . . . 48

3.2 Shell particle shapes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49


3.2.1 Flexible shells . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

3.3 Solid particle shapes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51


3.3.1 Flexible solid particles . . . . . . . . . . . . . . . . . . . . . . . 52
3.3.2 Convex and concave solid particles . . . . . . . . . . . . . . . . 55

4 Collision statistics 57

4.1 Collision statistics types . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58


4.1.1 Event-based statistics . . . . . . . . . . . . . . . . . . . . . . . . 58
4.1.2 Integral-based statistics . . . . . . . . . . . . . . . . . . . . . . . 60
4.1.2.1 Works . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.1.2.2 Impulses . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.1.2.3 Frequency . . . . . . . . . . . . . . . . . . . . . . . . 63

4.2 Collision statistics modules . . . . . . . . . . . . . . . . . . . . . . . . . . 63


4.2.1 Boundary collision statistics . . . . . . . . . . . . . . . . . . . . 63
4.2.1.1 Curves . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.2.2 Intra-particle collision statistics . . . . . . . . . . . . . . . . . . 66
4.2.3 Inter-particle collision statistics . . . . . . . . . . . . . . . . . . 69
4.2.4 Inter-group collision statistics . . . . . . . . . . . . . . . . . . . 70

5 Miscellaneous topics 72

5.1 Contact detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

5.2 Timestep calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76


5.2.1 Timestep for the hysteretic linear spring model . . . . . . . . . 77
5.2.2 Timestep for the linear spring-dashpot model . . . . . . . . . . 77
5.2.3 Timestep for the Hertzian spring-dashpot model . . . . . . . . 77

© 2020, esss - all rights reserved


dem technical manual iv
ROCKY

5.2.4 Numerical softening factor . . . . . . . . . . . . . . . . . . . . . 78

5.3 Sieve size calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

6 Bibliography 81

© 2020, esss - all rights reserved


dem technical manual 1
ROCKY

This document provides a brief introduction to the Discrete


Element Method (DEM) used in Rocky, and provides technical
details, equations, and further resources for understanding the various
particle simulation models used within it. Used in conjunction with
other Rocky documentation resources, including the User Manual and
the CFD Coupling Manual, this DEM Technical Manual should help
provide Rocky users with a more complete picture of how Rocky
works.

1 Introduction to DEM
Discrete Element Method (DEM) is a numerical technique for
predicting the behavior of bulk solids. A bulk solid is a large
collection of solid particles; otherwise known as granular media.
Some examples of granular media flows include grains being moved
through processing equipment, ore being passed through mining
machinery, and sand falling through an hourglass. Granular media
flow can be quite complex as these flows are known to exhibit solid-
like, fluid-like, or a combination of both behaviors. For example, sand
in an hourglass behaves like a fluid while a stockpile of sand can have
a solid-like stress-strain response.
DEM is a mesh-free method and does not solve the continuum
equations of motion. Hence, no stress-strain constitutive law for
the material is needed. Instead, a stress-strain relationship can
be obtained as an output from the DEM model. A general DEM
algorithm is shown in Figure 1.1.
The equations of motion for every individual particle are numer-
ically integrated with time. For this process, the total force on a
particle needs to be known. The total force is the resultant of contact
forces (between particles and with the boundary) and body forces.
Typical body forces are gravity (weight), fluid and other forces, such
as electrostatic, electromagnetic, and so on.

© 2020, esss - all rights reserved


dem technical manual 2
ROCKY

SETUP: User imports geometries, sets up particle groups, and


determines particle-to-particle and particle-to-boundary interactions
for the simulation project.

PROCESS: User chooses to begin processing the simulation. For each


individual particle, the DEM program does the following:
• Locates all neighboring particles and boundaries with which the
particle will come into contact.
• Calculates the sum of all forces and moments (Euler equations)
acting upon the particle: ‡
dv
∑ Fnet = ∑ Fbody + ∑ Fsurface = m dt

MOVE: The DEM program uses the current particle position, velocity,
and timestep information to move the particle to its next location in
the simulation: Z t+∆t
∑ Fnet
vnew = vold + dt
t m
Z t+∆t
xnew = xold + vnew dt
t

REPEAT:
Simulation end
time not reached and
YES
particles still
in range?

NO

Figure 1.1: Schematic algorithm of a


STOP
DEM program. As usual, F is force, m
is mass, t is time, v is velocity and x
is position. ‡ The Euler equations of
motion are not shown in the diagram
for simplicity.

© 2020, esss - all rights reserved


dem technical manual 3
ROCKY

2 Physical models in Rocky


Rocky is a software package developed by ESSS Rocky, S.L. that is
evolving very actively, with many new models implemented on a near
monthly basis. This chapter is a general guide for most of the models
currently implemented in Rocky. For specific information about the
fluid coupling models included in Rocky, please see the CFD Coupling
Technical Manual.

2.1 Contact force models


The contact forces in any DEM code (including Rocky) consist of the
following two parts:

• Forces normal-to-contact plane


• Forces tangent-to-contact plane

For spherical particles, the contact plane is perpendicular to the


line that connects the centers of two spheres. In the case of particle-
to-boundary contact, the line connects the center of a sphere and
the closest point of a triangle making up a boundary. For non-
round particles, the algorithm for determining a contact plane is more
complex; it involves calculating one of any of the following:

• The closest points of two particles.


• The closest points of a particle and triangle.
• The two points with the maximum overlap distance in the case
of a physical contact.

The contact plane is a plane perpendicular to the line connecting


these closest points. The descriptions for the normal and tangential
force calculations used in Rocky, as well as the models used by these
forces are presented below.

© 2020, esss - all rights reserved


dem technical manual 4
ROCKY

2.1.1 Normal force models

A normal force model for a DEM simulation has two major


requirements. First, the force has to be a repulsive one. Second, the
normal contact force model has to allow significant energy dissipation
since a granular medium is an extremely dissipative system. A
number of models have been proposed for these purposes. The
models implemented in Rocky are discussed below.

2.1.1.1 Hysteretic linear spring model

This model, first proposed by Walton & Braun1 , was referred to as 1


Walton, O. R. and Braun, R. L. (1986).
Viscosity, granular-temperature, and
linear hysteresis model in previous versions of Rocky. This elastic-plastic stress calculations for shearing assem-
(repulsive and dissipative) normal contact model allows simulation blies of inelastic, frictional disks. Journal
of Rheology, 30:948–980
of the plastic energy dissipation on a contact without introducing the
overhead of long simulation times.
In addition, since no viscous damping term is used, the energy
dissipation is not dependent on the relative velocities of neighboring
particles, making the energy dissipation insensitive to other contacts.
An additional advantage of this model is that compressible materials
can be accurately modeled due to the fact that the contact forces can
be almost zero even at residual overlaps.
The hysteretic linear spring model is implemented in Rocky in an
incremental way, as described by the following set of equations:

 min K st , F t−∆t + Knu ∆sn  if ∆sn > 0
t nl n n
Fn = (2.1)
 max F t−∆t + K ∆s , λ K st  if ∆sn < 0
n nu n nl n

∆sn = stn − stn−∆t (2.2)

where:

• Fnt and Fnt−∆t are the normal elastic-plastic contact forces at the
current time t and at the previous time t − ∆t, respectively, where
∆t is the timestep.

• ∆sn is the change in the contact normal overlap during the current
time. It is assumed to be positive as particles approach each other
and negative when they separate. stn and stn−∆t are the normal
overlap values at the current and at the previous time, respectively.

• Knl and Knu are the values of loading and unloading contact
stiffnesses, respectively.

© 2020, esss - all rights reserved


dem technical manual 5
ROCKY

• λ is a dimensionless small constant. Its value in Rocky is 0.001.


The part of the expression in which this constant is active ensures
that, during the unloading, the normal force will return to zero
when the overlap decreases to zero.

A typical cycle of loading/unloading is depicted in Figure 2.1.


Between points A and B, we have the loading process, in which
the normal force overlap increases linearly with slope Knl . After
reaching the maximum overlap, the unloading follows a steeper line
between points B and C, in which the slope is Knu . The plastic
deformation for the contact only exists during the contact, so any
residual deformation is forgotten after the contact ceases. The energy
dissipated in the collision is numerically equal to the shaded area in
the force–deformation diagram of Figure 2.1.

B From A to B

ing
ad
Lo From B to C

From C to A
A
C

Figure 2.1: Schematics of a typical


normal force–overlap response for the
hysteretic linear spring model.
The loading and unloading stiffnesses are defined by the particle
size, the bulk Young’s modulus, and by the restitution coefficient
of contacting materials, the last two of which the user inputs into
Rocky. The coefficient of restitution ε in Rocky is a measure of energy
dissipation for the contacting pair of materials. For the contact of two
particles, or of a particle with a boundary, the loading and unloading
equivalent stiffnesses are defined, respectively, as:

1 1


Knl,p 1 + Knl,p 2 for particle-particle contact
1 
= (2.3)
Knl  1 1

 Knl,p + Knl,b for particle-boundary contact

Knl
Knu = (2.4)
ε2

© 2020, esss - all rights reserved


dem technical manual 6
ROCKY

where subscripts 1 and 2 refer to particle 1 and particle 2 of two


contacting particles. The individual stiffnesses associated to a particle
and to a boundary are computed, respectively, as:

Knl,p = E p L (2.5)

Knl,b = Eb L (2.6)
where:

• E p is the particle material’s bulk Young’s modulus or elastic


modulus, which the user can input in the corresponding material’s
editor panel.2 2
A value of 1.0 × 108 Pa for the particle
bulk Young’s modulus is a reasonable
• Eb is the boundary material’s Young’s modulus, which is also a number for most particle sizes.

user input.

• L is the particle size.

In long-term contacts, for instance, among particles in a stockpile,


the hysteretic linear spring model can give rise to oscillations of very
small amplitudes on the normal force and on the overlap. Although
these oscillations are barely noticeable, this phenomenon can prevent
the particles from reaching a state of absolute repose. Because of
this, an additional mechanism of energy dissipation was introduced
in Rocky, in order to dissipate spurious oscillations in long-term
contacts. This mechanism consists in the addition of a viscous force
which is only activated during secondary loading cycles on long-term
contacts.3 That additional force is defined in a similar fashion to that 3
This force does not act during reg-
ular collisions, therefore, the energy
of the dissipative part of the contact force on the linear spring-dashpot dissipation specified by the restitution
model: coefficient will not be altered at all by
this force on those collisions.
Fn,v = Cn ṡn (2.7)

where ṡn is the time derivative of the normal overlap and Cn is the
damping coefficient, defined as in equation (2.9). The value of the
damping ratio η needed in that equation can be specified by the
user in the Advanced sub-tab of the Solver panel,4 in which it is listed 4
See also "About the Solver Parameters"
topic in the Rocky User Manual.
as Damping Ratio for Hysteretic Linear Spring Model . Allowed values for this
parameter are in the range from 0 to 1.5 5
The higher the value of the damping
ratio, the faster will be the dissipation
The hysteretic linear spring model works with three of the four of oscillations.
adhesive force models in Rocky, including the constant adhesive force
model, the linear adhesive force model and the Leeds adhesive force
model. Moreover, two models of tangential forces are compatible: the
linear spring Coulomb limit model and the Coulomb limit model.

© 2020, esss - all rights reserved


dem technical manual 7
ROCKY

2.1.1.2 Linear spring-dashpot model

The linear spring-dashpot model was referred to as linear elastic viscous


model in previous versions of Rocky. This phenomenological model,
first proposed in the seminal paper of Cundall & Strack6 , is widely 6
Cundall, P. A. and Strack, D. L. (1979).
A discrete numerical model for granular
used in DEM, mainly because of its simplicity. The normal contact assemblies. Geotechnique, 29(1):47–65
force in this model is composed of a linear elastic repulsive force and
a damping force, that is:

Fn = Knl sn + Cn ṡn (2.8)


where:

• Knl is the normal contact stiffness, defined previously for the linear
hysteretic spring model in equation (2.3).
• Cn is the normal damping coefficient.
• sn is the contact normal overlap.
• ṡn is the time derivative of the contact normal overlap.

The energy dissipation in the linear spring-dashpot model is


viscous in nature and due exclusively to the damping force term
in equation (2.8). The value of the normal damping coefficient Cn can
be determined in a way that the viscous energy dissipation matches
the energy dissipation of an inelastic collision, determined in turn
by the value of the coefficient of restitution. In order to do that, the
damping coefficient is defined in Rocky as follows:
p
Cn = 2 η m∗ Knl (2.9)
where:

• η is the damping ratio, a dimensionless parameter whose value is


related to the restitution coefficient (see below).
• m∗ is the effective mass for the contact, defined as:

1 1
1 
m1 + m2 for particle-particle contact
= (2.10)
m∗  1
for particle-boundary contact
m

where m1 and m2 are the masses of the contacting particles,


whereas m is the mass of the particle in contact with a boundary.

In general, there will be one value of the damping ratio η for which
the energy dissipation on a collision event modeled with the linear
spring-dashpot will replicate the energy dissipation predicted by the
coefficient of restitution.7 The functional relationship between the 7
The coefficient of restitution is consid-
ered in Rocky to be a property of the
damping ratio and the coefficient of restitution, derived from that interaction between two materials.
condition, is:

© 2020, esss - all rights reserved


dem technical manual 8
ROCKY

   √ 
2 η 1− η 2

 exp − √ η
π − arctan if 0 6 η < √1


 1− η 2 1−2η 2 2



  
2 η 1− η 2

ε= exp − √ 2 arctan 2η 2 −1
η
if √1 6η61 (2.11)
 1− η 2


  √ 
η + η 2 −1


 exp − 2

 √ η
ln √ 2
if η > 1
η −1 η− η −1

As can be seen in Figure 2.2, equation (2.11) defines the restitution


coefficient ε as a monotonic function of the damping ratio η. Since the
inverse of this function cannot be determined analytically, equation
(2.11) is solved numerically in Rocky in order to find the value of η
that corresponds to the value of ε prescribed by the user.

1
Figure 2.2: Graph of the relationship
between the damping ratio η and the
restitution coefficient ε, given by equa-
0.75 tion (2.11).

0.5

0.25

0 0.5 1 1.5 2 2.5 3

Equation (2.11) was derived originally by Schwager & Pöschel8 , 8


Schwager, T. and Pöschel, T. (2007).
Coefficient of restitution and linear-
considering that the end of a collision happens when the contact dashpot model revisited. Granular Mat-
normal force decreases to zero. It is common in DEM to see the use ter, 9:465–469

of a simpler expression for the relation between ε and η, derived


from the condition that the end of a collision occurs at a later time,
when the deformation returns to zero. It is easy to show, however,
that the normal force given by equation (2.8) turns negative at that
time. Since a negative normal contact force is an attractive one, the
assumption of a purely repulsive contact force is contradicted by that
practice. On the other hand, equation (2.11) is totally consistent with
that assumption and for that reason it was adopted in Rocky.
Although the linear spring-dashpot model is popular in DEM
formulations, it tends to be less accurate than the hysteretic linear
spring model because energy dissipation in the real world is plastic
rather than viscous. Its accuracy drops especially when particles

© 2020, esss - all rights reserved


dem technical manual 9
ROCKY

have simultaneous multiple contacts, since the amount of viscous


dissipation is accurate only for single contacts.
The linear spring-dashpot model works with the constant adhesive
force model and the linear adhesive force model, as well as the
linear spring Coulomb limit model and the Coulomb limit model for
tangential forces.

2.1.1.3 Hertzian spring-dashpot model

The Hertzian spring-dashpot model is new to Rocky as of v4.1.


This model is similar to the linear spring-dashpot model, the main
difference being that both the elastic and the damping components
of the normal force are nonlinear functions of the overlap in the
Hertzian model. The elastic part is based in the classical contact
theory developed by Hertz in the late nineteenth century.9 9
Hertz, H. (1882). Über die berührung
fester elastischer körper (On the contact
The form of the Hertzian spring-dashpot model implemented in of elastic solids). J. Reine Angewandte
Rocky can be written as: Mathematik, 92:156–171. English transla-
tion, Macmillan, London, 1896
3 1
Fn = K̂ H sn2 + ĈH sn4 ṡn (2.12)

in which the stiffness coefficient K̂ H is defined as:10 10


It is worth noting that K̂ H and ĈH
do not have the same units as the
√ corresponding coefficients in the linear
K̂ H = 34 E∗ R∗ (2.13) spring-dashpot model. This is because
where: of the different functional form of equa-
tions (2.8) and (2.12).

• E∗ is the effective Young’s modulus, defined by the expression

1 1 − ν12 1 − ν22

= + (2.14)
E E1 E2

in which E1 and E2 are the Young’s moduli of the two contacting


particles or the particle and the boundary, depending on the contact
type. Moreover, ν1 and ν2 are the respective Poisson’s ratios.

• R∗ is the effective or equivalent radius, defined by the expression:



2 2
1 
L1 + L2 for particle-particle contact
= (2.15)
R∗  2
for particle-boundary contact
L

in which L1 and L2 are the sizes of the contacting particles, whereas


L is the size of the particle in contact with the boundary.

Following the practice proposed initially by Tsuji et al.11 , the 11


Tsuji, Y., Tanaka, T., and Ishida, T.
(1992). Lagrangian numerical simulation
damping coefficient for the Hertzian model in Rocky is defined in a of plug flow of cohesionless particles
in a horizontal pipe. Powder Technology,
71:239–250

© 2020, esss - all rights reserved


dem technical manual 10
ROCKY

similar fashion as for the linear spring-dashpot model:


q
ĈH = 2 η H m∗ K̂ H (2.16)

where m∗ is the effective mass, defined previously in equation (2.10),


and η H is the damping ratio for the Hertzian spring-dashpot model.
Antypov & Elliott12 demonstrated that considering appropriate 12
Antypov, D. and Elliott, J. A. (2011).
On an analytical solution for the
variable transformations, the solution for a movement equation based damped hertzian spring. Europhysics
on the Hertzian model can be mapped onto the solution of an Letters, 94(5)

equivalent equation based on the linear spring-dashpot model. That


equivalence is possible only if the damping coefficient for the Hertzian
model is defined as: √
5
ηH = η (2.17)
2

where η is the same damping ratio used in the linear spring-dashpot


model. In Rocky, the value of η is determined solving equation (2.11)
and then using the resulting value in equation (2.17) to obtain η H .
This guarantees that no matter whether the linear spring-dashpot
or the Hertzian model is used, the viscous energy dissipation in
a collision event is equal to the one predicted by the coefficient of
restitution.
The Hertzian spring dashpot model works with all three models
available for tangential forces, namely, the linear spring Coulomb limit
model, the Coulomb limit model, and the Mindlin-Deresiewicz model.
Regarding the adhesive models, it only works with the constant
adhesive force model, the linear adhesive force model, and the JKR
adhesive force model.

2.1.2 Tangential force models

The models used in Rocky to calculate the tangential components of


the contact forces are described below.

2.1.2.1 Linear spring Coulomb limit model

This model was referred to as elastic Coulomb model in previous


versions of Rocky. The tangential force in this model is elastic-
frictional. If the tangential force were considered purely elastic, its
value at time t would be given by:13 13
The minus sign in this equation arises
from the fact that the tangential force
always opposes to the tangential dis-
Ftτ,e = Ftτ−∆t − Kτ ∆sτ (2.18) placement.

© 2020, esss - all rights reserved


dem technical manual 11
ROCKY

where:

• Ftτ−∆t is the value of the tangential force at the previous time.


• ∆sτ is the tangential relative displacement of the particles during
the timestep.
• Kτ is the tangential stiffness defined as:

Kτ = rK Knl (2.19)

where Knl is the loading normal stiffness defined in equation (2.3)


and rK is a user-defined parameter listed as Tangential Stiffness Ratio
in the corresponding materials interaction editor panel.

In this model, however, the tangential force cannot exceed the


Coulomb’s limit. Therefore, the complete expression for the tangential
force is:
  Ft
Ftτ = min |Ftτ,e |, µFnt
τ,e
(2.20)
|Ftτ,e |
where:

• Fnt is the contact normal force at time t.


• µ is the friction coefficient, defined as:

 µs if no sliding takes place at the contact
µ= (2.21)
 µ if sliding does take place at the contact
d

in which µs and µd are, respectively, the static and the dynamic


friction coefficients. These are properties that the user can define in
the corresponding materials interaction editor panel in Rocky.
The sliding is considered to be taking place on the contact the first
time the tangential force Ftτ,e exceeds the limit of µs Fnt . Once that force
falls below the value of µs Fnt , the contact is considered non-sliding
again.
This tangential model works with both the hysteretic linear spring
model and the linear spring-dashpot model for normal forces.

2.1.2.2 Coulomb limit model

This is the simplest tangential force model implemented in Rocky.


The tangential force according to this model is given by:

ṡτ
Fτ = −µ Fn (2.22)
|ṡτ |
where:

• µ is the friction coefficient defined in equation (2.21).

© 2020, esss - all rights reserved


dem technical manual 12
ROCKY

• Fn is the normal force at the contact.


• ṡτ is the tangential component of the relative velocity vector.

The condition of sliding considered for this model is based on the


tangential component of the relative velocity. If that component is
below vτ,lim , the contact is considered non-sliding. In the current
version of Rocky, the value of vτ,lim is set to 0.001 m/s.
The Coulomb limit model works with all three types of normal
forces models, i.e., the hysteretic linear spring model, the linear
spring-dashpot model and the Hertzian spring-dashpot model.

2.1.2.3 Mindlin-Deresiewicz model

This model is new to Rocky as of v4.1. The tangential force in this


model is given by the expressions:
s
  s
3 τ 6 µm∗ Fn 1
Fτ = −µ Fn 1−ς 2 + ητ ς 4 ṡτ (2.23)
|sτ | sτ, max

min(|sτ |, sτ, max )


ς = 1− (2.24)
sτ, max
where:

• µ is the friction coefficient, defined in equation (2.21).


• Fn is the normal force.
• sτ is the tangential relative displacement at the contact.
• ṡτ is the tangential component of the relative velocity at the contact.
• sτ, max is the maximum relative tangential displacement at which
particles begin to slide.
• m∗ is the effective mass, defined in equation (2.10).
• ητ is the tangential damping ratio, which is estimated in Rocky by
means of:
ln ε
ητ = − p (2.25)
2
ln ε + π 2

where ε is the coefficient of restitution for the respective materials


interaction.

The value of the maximum relative tangential displacement sτ, max


is determined in Rocky using the expression:
−1
1 − ν1 1 − ν2

sτ, max = µ + sn (2.26)
2 − ν1 2 − ν2
where:

© 2020, esss - all rights reserved


dem technical manual 13
ROCKY

• ν1 and ν2 are the Poisson’s ratios of the two particles or the particle
and the boundary, depending on the contact type.
• sn is the normal overlap.

It is worth noting that when |sτ | > sτ, max , the parameter ς turns
to zero in equation (2.24). It is easy to show that the magnitude of
the tangential force reduces to µ Fn in that case, and, therefore, the
Mindlin-Deresiecz model becomes equivalent to the Coulomb limit
model under that specific condition.
The Mindlin-Deresiewicz model works only with the Hertzian
spring-dashpot model. Moreover, the Mindlin-Deresiewicz model
does not work properly with rolling resistance models for non-round
particles. Therefore, when this model is selected in a problem with
non-round particles, the rolling resistance must be set to zero.

2.2 Adhesive force models


The elastic-plastic normal contact model is appropriate for simulating
non-adhesive and dry granular materials. Unfortunately, quite often
granular materials can have various moisture contents that can cause
them to adhere to themselves and other surfaces with which they
come into contact. In order to capture this behavior, the repulsive
normal force has to be supplemented by the attractive normal force
to accurately predict its flow characteristics. For example, this is
required for wet and sticky materials where the adhesion is from
liquid bridge forces or other mechanisms.
One important point to note is that cohesion/adhesion of a bulk
solid is a function of the stress. The linear force adhesion contact
model in Rocky captures this physical phenomenon by scaling the
cohesion/adhesion with the force of contact.
Referred to as adhesion models in previous versions of Rocky, as
of v4.1 the now called adhesive force models are set in the same UI
location as the other models, and are described in the sections below.

2.2.1 Constant adhesive force model

With proper calibration the constant adhesive force model can be used
to simulate the behavior of adhesive materials that do not exhibit
stress consolidation effects, e.g., when adhesion is due to liquid bridge
forces.

© 2020, esss - all rights reserved


dem technical manual 14
ROCKY

This is the simplest adhesion model in Rocky and is described by


the equation:

 0 if − sn > δadh
Fn,adh = (2.27)
g min(m1 , m2 ) if − sn < δadh
 f
adh

where:

• Fn,adh is the normal adhesive contact force.

• sn is the contact normal overlap.14 14


The normal overlap is assumed to be
positive as particles approach each other
• m1 and m2 are the mass of the particles in contact. and negative when they separate.

• g is the gravity acceleration.

• δadh is a model parameter listed as Adhesive Distance in the Rocky


UI. If the distance between two particles or between a particle
and a boundary surface is below δadh , the adhesive force will be
activated. In Rocky, the value of this parameter can be defined in
the materials interaction editor panel.

• f adh is a model parameter listed in the Rocky UI as Force Fraction .


The value of the adhesive force will be equal to the particle gravity
force multiplied by the value of f adh . If the force fraction is 1, that
means the adhesive force will be equal to the gravity force on the
particle. In the case of contact between two particles of different
mass, the smallest mass is considered for gravity force calculations.

The constant adhesive force model works with all three types of
normal force models in Rocky, including the hysteretic linear spring
model, the linear spring-dashpot model and the Hertzian spring-
dashpot model.

2.2.2 Linear adhesive force model

Like the constant adhesive force model, the linear adhesive force
model requires calibration of the material behavior. This model is
appropriate for granular materials with stress consolidation.
The linear adhesive model, in essence, behaves like an attractive
linear spring; the value of the adhesive force is zero if the contact
distance is larger than the minimum adhesive contact force distance,
and increases proportionally to the difference between this distance
and the actual particles’ contact distance/overlap. The coefficient of
proportionality is defined as the adhesive force stiffness.

© 2020, esss - all rights reserved


dem technical manual 15
ROCKY

The linear adhesive force model as implemented in Rocky can be


described by the following expression:

0 if − sn > δadh
Fn,adh = (2.28)
r
adh Knl (sn + δadh ) otherwise
where:

• Fn,adh is the normal adhesive contact force at the current time.

• sn is the contact normal overlap during the current time (it is


assumed to be positive as particles approach each other and
negative when they separate).

• δadh is a model parameter listed as Adhesive Distance in the Rocky


UI. If the distance between particles or between a particle and a
boundary surface is less than this parameter, the adhesive force
will be activated.

• r adh is a model parameter listed as Stiffness Fraction in the Rocky UI.


This parameter is defined as the ratio of the adhesive force stiffness
to the contact loading stiffness.

• Knl is the loading contact normal stiffness.

Figure 2.3 shows a force-overlap plot of a particle collision with a


wall for both dry and adhesive impacts. It is evident that the larger
the contact overlap, the larger the attractive force, thus allowing
an accurate model to represent the stress consolidation of granular
materials.

no adhesion stiffness fraction = 0.05 stiffness fraction = 0.25 stiffness fraction = 0.5
Figure 2.3: Examples of force–overlap
plots for different adhesive stiffness
fractions.
25

20

15

10
Total force [N]

0
0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035 0.004

-5

-10

-15

Overlap/diameter

© 2020, esss - all rights reserved


dem technical manual 16
ROCKY

The linear adhesive model works with all three types of normal
force models in Rocky, including the hysteretic linear spring model,
the linear spring-dashpot model and the Hertzian spring-dashpot
model.

2.2.3 Leeds adhesive force model

New to Rocky v4.1, the Leeds adhesive force model is a simplified


linear contact model that simulates elasto-plastic-adhesive behavior
based upon the model developed by Pasha, from the University of
Leeds.15 , 16 15
Pasha, M. (2013). Modelling of flowabil-
ity measurement of cohesive powders using
Figure 2.4 shows schematically a typical normal force–overlap small quantities. PhD thesis, University
response of the Leeds model, which is determined by the piecewise of Leeds, UK
16
Pasha, M., Dogbe, S., Hare, C., Has-
function:
sanpour, A., and Ghadiri, M. (2014).
 A linear model of elasto-plastic and


 −Knu sn + 10
9 Fce from A → B adhesive contact deformation. Granular
Matter, 16:151–162


K s + 8F

from B → C

 nu n 9 ce



Fn = Knl (sn − so ) from C → D (2.29)



Knu (sn − s p ) from D → E






 −K (s + s − 2 s )

from E → F
nu n p cp

D Figure 2.4: Schematic diagram of a


typical normal force–overlap response
of the Leeds model.

F
B

© 2020, esss - all rights reserved


dem technical manual 17
ROCKY

In equation (2.29), Knl and Knu are the loading and unloading
stiffnesses, defined in equations (2.3) and (2.4), respectively. Moreover,
Fce is the elastic pull-off force predicted by the JKR theory, which is
given by:
Fce = − 23 πR∗ Γ (2.30)

where:

• R∗ is the effective particle radius, defined previously in equation


(2.15).

• Γ is a model parameter listed as Surface Energy in the corresponding


materials interaction editor panel.

The overlap values that delimit the different stages in equation


(2.29) are given by:  
Fce
so = − 89 (2.31)
Knu
Knl
s p = smax − (smax − so ) (2.32)
Knu

q
162 πΓ
scp = s p − 137 Knu (s p − so )(2R∗ − s p + so ) (2.33)

where smax is the maximum overlap reached, as can be seen in Figure


2.4.
The Leeds model is similar to the JKR model, since both introduce
the concept of surface energy to model adhesion. The main difference
is that the Leeds model was adapted to the hysteretic linear spring
model, whereas the JKR model was developed originally as an
extension of the Hertzian contact model. The main attraction for both
models is that the surface energy can be measured experimentally for
well-defined cases. Therefore, in principle, if an experimental value
for that parameter were available and real elastic properties were
used, simulations could be performed without any calibration.
The main disadvantage of the Leeds model, compared to the JKR
model, is that the adhesive force is based on usually high values of
unloading stiffness. The proper treatment of this situation usually
will require timesteps lower than the default values used in Rocky
and, consequently, simulations usually will run slower.
The Leeds adhesive model works only with the hysteretic linear
spring normal contact force model.

© 2020, esss - all rights reserved


dem technical manual 18
ROCKY

2.2.4 JKR adhesive force model

The Johnson-Kendall-Roberts17 , JKR, model introduces the effect of 17


Johnson, K. L., Kendal, K., and
Roberts, A. D. (1971). Surface energy
adhesion into the Hertzian contact model. In the JKR model, the and the contact of elastic solids. Proceed-
contact area between two particles is slightly larger than the one ings of the Royal Society A: Mathematical,
Physical and Engineering Sciences, 324:301–
predicted by the Hertzian theory. This is the result of a surface energy 313
term that is added to the total energy of the system.
The adhesive part of the normal force when the JKR model is
employed is given by:18 18
Johnson, K. L. (1985). Contact mechan-
√ ics. Cambridge University Press
Fn,adh = 8πΓE∗ a3 (2.34)
where:

• Γ is a model parameter listed as Surface Energy in the Rocky UI. The


value of this parameter can be introduced in a materials interaction
editor panel.
• E∗ is the effective Young’s modulus, defined in equation (2.14).
• a is the radius of contact between particles or between a particle
and a boundary.

The contact radius is related to the normal overlap sn by means of


the expression: 1/2
a2

2πΓa
sn = ∗ − (2.35)
R E∗

where R∗ is the effective particle radius, defined in equation (2.15).


For a given normal overlap sn , equation (2.35) is solved in Rocky
to determine the corresponding value of a, in order to calculate the
adhesive force using equation (2.34).
When the JKR model is enabled, the elastic part of the Hertzian
normal force must be corrected, because the contact radius in the JKR
model is different from the radius predicted by the Hertzian model.
The corrected elastic force, expressed as a function of the radius of
contact a, is given by:
4E∗ a3
Fn,e = (2.36)
3R∗

The JKR model has a strong theoretical basis and is widely accepted
for adhesive elastic spheres. Since the surface energy can be measured
experimentally, this model could be used without calibration, in
principle, for simulating perfect spheres. In any other case, calibration
still will be needed. In Rocky, the JKR adhesive force model only
works with the Hertzian spring-dashpot model.

© 2020, esss - all rights reserved


dem technical manual 19
ROCKY

2.3 Rolling resistance models


Rolling resistance is the common name used when a moment that
opposes the rolling motion of a particle is introduced into the
modeling. This moment is usually incorporated as a practical way
to represent the effect of non-sphericity on rolling spheres or the
effect of surface irregularities on other type of particles. Two rolling
resistance models are available in Rocky, named Type 1 and Type 3 .
These models correspond to models type A and type C, respectively,
in the classification proposed by Ai et al.19 19
Ai, J., Chen, J. F., Rotter, J. M., and
Ooi, J. Y. (2010). Assessment of rolling
resistance models in discrete element
simulations. Powder Technology, 206:269–
2.3.1 Rolling resistance type 1 282

In this model, a constant moment is applied to the particle in order


to represent rolling resistance. The mathematical expression for this
moment is:
ω
Mr = −µr |r| Fn (2.37)
|ω |
where:

• µr is the rolling resistance coefficient, defined in Rocky as a particle


property listed simply as Rolling Resistance . Ai et al.20 define this 20
Ai, J., Chen, J. F., Rotter, J. M., and
Ooi, J. Y. (2010). Assessment of rolling
dimensionless parameter as the tangent of the maximum angle of resistance models in discrete element
a slope on which the rolling resistance moment counterbalances simulations. Powder Technology, 206:269–
282
the moment produced by gravity in the particle.
• Fn is the contact normal force.
• ω is the particle angular velocity vector. The direction of the rolling
resistance moment vector will coincide with the direction of this
angular velocity.
• |r| is the particle rolling radius, where r is the vector joining the
centroid of the particle and the contact point.

The rolling resistance model type 1 only should be used in


situations when a high angle of repose is needed without using
adhesive force models.

2.3.2 Rolling resistance type 3

This is an elastic-plastic model that is the recommended one for most


simulations in Rocky. This model usually includes a viscous damping
term, but as Wensrich & Katterfeld21 argue, the proper choice of 21
Wensrich, C. M. and Katterfeld, A.
(2012). Rolling friction as a technique
the rolling stiffness value provides a good behavior of the rolling for modelling particle shape in DEM.
Powder Technology, 217:409–417

© 2020, esss - all rights reserved


dem technical manual 20
ROCKY

resistance without any additional damping. This is the approach


followed in Rocky, where the rolling stiffness is defined as:

Kr = R2r Kτ (2.38)

where Kτ is the tangential stiffness, defined in equation (2.19),


whereas Rr is the rolling radius, given by:

1 1
1 
| r1 |
+ | r2 |
for particle-particle contact
= (2.39)
Rr  1
for particle-boundary contact
|r|

in which r1 and r2 are the rolling radii of the contacting particles,


whereas r is the rolling radius of the particle in contact with the
boundary. The rolling radius vector is defined as the vector joining
the centroid of the particle and the contact point at a given time.
If the rolling resistance were purely elastic, the rolling resistance
moment would be updated incrementally in the following way:22
22
The minus sign in equation (2.40)
guarantees that the rolling resistance
t
Mr,e = Mrt−∆t − Kr ω rel ∆t (2.40) moment always opposes the relative
rolling motion.
where:

• Mrt−∆t is the rolling resistance moment vector at the previous time.


• Kr is the rolling stiffness defined in equation (2.38).
• ω rel is the relative angular velocity vector, which is defined as the
difference between the angular velocities of two contacting particles
or the angular velocity of a particle on a boundary, as the case may
be.
• ∆t is the simulation timestep.

The updated rolling resistance moment defined in equation (2.40)


is not used directly in the movement equation for the particles. In
the rolling resistance model type 3, the magnitude of the rolling
resistance moment is limited by the value which is achieved at a full
mobilization rolling angle. The limiting value is:

Mr,lim = µr Rr Fn (2.41)
where:

• µr is the rolling resistance coefficient, defined in Rocky as a particle


property listed simply as Rolling Resistance . As stated previously,
this dimensionless parameter is defined as the tangent of the
maximum angle of a slope on which the rolling resistance moment
counterbalances the moment produced by gravity in the particle.

© 2020, esss - all rights reserved


dem technical manual 21
ROCKY

• Rr is the rolling radius defined in equation (2.39)


• Fn is the contact normal force.

The final expression for the rolling resistance moment in model


type 3 is:   Mt
r,e
Mrt = min |Mr,e
t
|, Mr,lim t |
(2.42)
|Mr,e

2.4 Breakage models


Rocky includes models for predicting the instantaneous breakage of
particles based upon the stressing energy involved in collisions with
walls or other particles. Those models work only with polyhedral
convex particles; therefore, no spherical, rounded or concave shapes
are able to break in Rocky using these models. When a particle breaks,
the resulting fragments preserve both mass and volume.

2.4.1 Ab-T10 breakage model


23
Vogel, L. and Peukert, W. (2005). From
The Ab-T10 breakage probability is based on the model posed by single particle impact behaviour to mod-
elling of impact mills. Chemical Engineer-
Vogel & Peukert23 and the subsequent modification proposed by Shi ing Science, 60(18):5164–5176
& Kojovic24 . This breakage model treats every particle as a single 24
Shi, F. and Kojovic, T. (2007). Vali-
dation of a model for impact breakage
entity that can be broken into fragments instantaneously based upon incorporating particle size effect. In-
the impact energy it receives. ternational Journal of Mineral Processing,
82(3):156–163
In Rocky, the total specific contact energy, ec , is computed by
summing the work done by the contact forces at all contact points
in a particle during the loading period.25 In order to damage the 25
The contact energy computed includes
both the elastic and the dissipated en-
particle, ec should be greater than the minimum breakage energy of ergy in the normal direction. Depending
the particle, emin . This minimum breakage energy is related to the on the normal contact model employed,
the dissipated energy will be either
particle size through the expression: plastic or viscous. In addition, the elastic
part of the energy transferred in the
Lref tangential direction is also considered
emin = emin,ref (2.43) in the computation.
L
where:

• emin,ref is the reference minimum specific energy value for a


reference particle size of this material. This parameter is referred
to as Reference Minimum Specific Energy in the Rocky UI.

• Lref is the reference particle size, referred to as Reference Size in the


Rocky UI.

• L is the actual particle size.

© 2020, esss - all rights reserved


dem technical manual 22
ROCKY

In order to take into account the damage caused by successive


collisions, the verification of breakage is made by considering a
cumulative value of the specific contact energy, ecum . If ect is the
instantaneous value of the specific contact energy at a given time
t during a loading phase, the value of ecum will be updated only if
ect > emin and ect > emax , where emax is the maximum value of contact
energy registered on the particle until the last time in which the value
of ecum was updated. When those conditions are satisfied, the update
is made according to the expression:
 
ecum = ecum + ect − max ect−∆t , emax (2.44)

where ect−∆t is the specific contact energy at the previous timestep.


Whenever the particle is unloaded and the value of ect decreases below
emin , the value of ec, max is reset to zero, so a new cycle of loading may
begin, in which the value of ecum will be able to increase again.
In the Ab-T10 model, the breakage probability for a given
cumulative specific contact energy value is calculated as:
 
P(ecum ) = 1 − exp −S ecum L/Lref (2.45)
where:

• S is a material constant referred to as Selection Function Coefficient in


the Rocky UI.

A particle will break if at any moment the value of P(ecum )


computed with equation (2.45) is larger than the strength of the
particle. When that condition is met and the particle breaks, the
fragments are generated following the Voronoi fracture algorithm
according to a size distribution that the user specifies. Two options are
currently available: Gaudin-Schumann and incomplete beta function.
For either option, the value of the necessary t10 parameter is calculated
according to the expression:
h  i
t10 = M 1 − exp −S ecum L/Lref (2.46)
where:

• t10 is the percentage of fragments passing a screen size of 1/10th


of the original particle size L.
• M is the maximum t10 for a material subject to breakage, referred
to as Maximum t10 Value in the Rocky UI.

New fragments generated by breakage can break further if they are


subjected to additional damage, generating even smaller fragments.

© 2020, esss - all rights reserved


dem technical manual 23
ROCKY

In order to prevent existing contacts at the moment of the breakage


to have an abnormal influence on the subsequent breakage of newly
generated fragments, the update of ecum using equation (2.44) is
skipped during a certain number of timesteps after a breakage event.
Currently, that number of timesteps is set to 25 in Rocky.
For more information about the Ab-T10 breakage model, please
refer to the papers of Potapov & Donahue26 and Shi & Kojovic27 . 26
Potapov, A. and Donahue, T. (2012).
Computer simulation of coal breakage
in conveyor transfer chutes with Rocky
discrete element method package. Tech-
2.4.2 Tavares breakage model nical report, Rocky DEM, Inc
27
Shi, F. and Kojovic, T. (2007). Vali-
This model is based on the PhD work of Prof. L. M. Tavares at the dation of a model for impact breakage
incorporating particle size effect. In-
University of Utah and the further development with his research ternational Journal of Mineral Processing,
group at the Federal University of Rio de Janeiro, Brazil. The Tavares 82(3):156–163

model extends the functionality of simpler breakage models by adding


capabilities that can make breakage prediction quite realistic in a wide
variety of situations. In particular, the model is useful in describing
ore degradation during handling as well as size reduction in different
types of crushers, providing greater confidence in predicting both the
proportion of broken particles and product size distribution.
In the Tavares breakage model, the breakage probability is based
on an upper-truncated log-normal distribution of the specific fracture
energy, e. This distribution is defined by the expression:28 28
Tavares, L. M. (2009). Analysis of
particle fracture by repeated stressing
ln e∗ − ln e50 as damage accumulation. Powder Tech-
  
1 √
P0 (e) = 2 1 + erf (2.47) nology, 90(3):327–339
2σ2
where:

• e∗ is the relative specific fracture energy, defined below in equation


(2.48).

• e50 is the median particle specific fracture energy.

• σ2 is the variance of the log-normal distribution of fracture energies.


This is a model parameter that must be specified by the user, in the
Rocky UI it is listed as σ2 .

The relative specific fracture energy is defined as:


emax e
e∗ = (2.48)
emax − e

where emax corresponds to the specific impact energy above which


all particles would break in a single impact. In the Rocky UI, the
value of emax is specified indirectly through the ratio emax /e50 , listed
as e max/e 50 .

© 2020, esss - all rights reserved


dem technical manual 24
ROCKY

The particle specific fracture energy is highly dependent on the


particle size, L. This dependency is introduced in the model by means
of the following correlation between the median specific fracture
energy, e50 , and the particle size:29 29
Tavares, L. M. and Carvalho, R. M.
  ϕ  (2009). Modeling breakage rates of
d0 coarse particles in ball mills. Mineral
e50 = e∞ 1 + (2.49) Engineering, 22:650–659
L

where e∞ , d0 , and ϕ are model parameters that should be fitted to


experimental data. These parameters are listed in the Rocky UI as
e∞ , d0 , and ϕ , respectively.
Whenever a particle enters the solution domain in Rocky, a random
strength is assigned to it. This property can be interpreted as the value
of P0 at which it will break during a simulation. Then, considering
equations (2.47)–(2.49), it can be determined the corresponding value
of e, the specific fracture energy of the particle. Here, this specific
fracture energy will be denoted as e0 , because this value is considered
to determine if the particle will break during the first collision
event. The exact criterion considered for the particle breakage will be
described below.
Every time a particle undergoes a collision event, the fracture
specific energy will decrease due to the accumulated damage to
the particle during the loading process. Therefore, after every new
loading cycle without breakage, a new particle specific fracture energy
is computed based on the previous one and on an estimation of
the accumulated damage during the loading. The mathematical
expressions considered for this are:30 30
Tavares, L. M. (2009). Analysis of
particle fracture by repeated stressing
as damage accumulation. Powder Tech-
en = en−1 (1 − Dn∗ ) (2.50) nology, 90(3):327–339

  2γ
2γ ec,n 5
Dn∗ = (2.51)
2γ − 5Dn∗ + 5 en−1
where:

• en is the particle specific fracture energy after n loading cycles


without breakage. Therefore, en−1 may be interpreted as the value
of en computed at the end of the previous loading cycle.
• Dn∗ is the fractional damage in the particle during the nth loading
cycle.
• γ is the damage accumulation coefficient, a model parameter listed
as γ in the Rocky UI.
• ec,n is the instantaneous specific contact energy in the particle at
the end of the nth loading cycle.

© 2020, esss - all rights reserved


dem technical manual 25
ROCKY

Let’s define ect as the instantaneous specific contact energy in the


particle at a given time t. In Rocky, that instantaneous contact energy
is computed by taking into account the work done by all contact
forces acting on the particle at that time. This includes the work done
by elastic and dissipative normal forces, as well as the work done by
elastic tangential forces.
Now, let’s consider that a particle is in its nth loading cycle.
Whenever the value of ect is greater than en−1 , that is, the specific
fracture energy at the end of the previous loading cycle,31 the particle 31
Or the initial specific fracture energy,
e0 , if it is the first loading cycle.
will break. If the unloading of the particle begins before that breakage
condition is satisfied, a new value of specific fracture energy may be
computed with equations (2.50) and (2.51). This new threshold value
for breakage in a subsequent loading cycle is calculated whenever ect
decreases in the unloading below emin , which is the minimum specific
energy for breakage. This is a user input parameter, listed in the
Rocky UI as e min .
When a particle breaks, the geometries of the resulting fragments
are generated by means of the Voronoi fracture algorithm, according
to a size distribution specified by the user between the two options
available: Gaudin-Schumann and incomplete beta function. The
value of the t10 parameter, needed in either of those distributions, is
calculated according to the expression:32 32
Tavares, L. M. (2009). Analysis of
particle fracture by repeated stressing
as damage accumulation. Powder Tech-
  
ec,b
t10 = A 1 − exp −b0 (2.52) nology, 90(3):327–339
êb
where:

• t10 is the percentage of fragments passing a screen size of 1/10th


of the original particle size L.

• A and b0 are model parameters listed as A and b’ , respectively,


in the Rocky UI.

• ec,b is the value of the specific contact energy in the particle at the
instant of breakage.

• êb is a measure of the specific fracture energy of the broken


particles.

Regarding the specific fracture energy êb in equation (2.52), two


alternatives are currently available in Rocky. The first one is to use the
value of the current specific contact energy at the instant of breakage,
ec,b . As this term will cancel out with the numerator in equation
(2.52), with this option the value of t10 will become independent of
the contact specific energy of the broken particle. In the Rocky UI,

© 2020, esss - all rights reserved


dem technical manual 26
ROCKY

this option is provided via the Energy for t10 Calculation parameter, and
is listed as Current Particle Energy (the default option).
The second option available is to use the value of the median
specific fracture energy of the broken particles e50b , defined as:33 33
Tavares, L. M. (2009). Analysis of
particle fracture by repeated stressing
√ h i as damage accumulation. Powder Tech-
e50b = e50 exp 2 σ2 erf−1 P0 (ec,b ) − 1 (2.53) nology, 90(3):327–339

This option is listed as Median Specific Fracture Energy (e50b) in the


Rocky UI.
As was the case with the Ab-T10 model, some preventive measures
are also taken in order to avoid excessive re-breakage of fragments
resulting from previous breakage events. First, in order to rule out
any influence of existing contacts on the breakage of new fragments,
calculations of damage are skipped during 25 timesteps after the
breakage event that produced them.34 Second, in order to reduce the 34
This value is not editable by the
user, since it was verified that it works
available energy for damaging newly formed fragments, a parameter satisfactorily in several scenarios.
listed as Fragments Energy Factor was introduced into the modeling.
The value defined for this parameter will multiply the value of
the instantaneous specific contact energy35 when considering the 35
Suitable values for this energy factor
are in the interval [0,1]
breakage criterion for fragments.
For more information about the Tavares breakage model, please
refer to the papers previously cited and also to Tavares & King
(1998)36 , Tavares & King (2002)37 , and Carvalho & Tavares (2013)38 . 36
Tavares, L. and King, R. (1998). Single-
particle fracture under impact loading.
International Journal of Mineral Processing,
54:1–28
2.4.3 Breakage for large deformations 37
Tavares, L. M. and King, R. P. (2002).
Modeling of particle fracture by re-
Besides the mechanisms of breakage based on fracture energy peated impacts using continuum dam-
age mechanics. Powder Technology,
described in previous sections, particles may also break in Rocky 123(2):138–146
when its deformation exceeds a certain preset value. This additional 38
Carvalho, R. M. and Tavares, L. M.
(2013). Predicting the effect of operating
mechanism is introduced in order to prevent some artifacts related and design variables on breakage rates
to low-stiffness particles that can experience large overlaps during a using the mechanistic ball mill model.
Minerals and Engineering, 43:91–101
simulation.
This additional breakage mechanism is activated by default in
Rocky, but users may disable it if necessary through the option
Set Breakage Overlap Factor , which is found in the Advanced sub-tab of
the Solver panel.39 When this option is enabled, users will be able to 39
The Advanced sub-tab is visible only
when the Advanced Features option is
set a value to the Breakage Overlap Factor . From that point on, whenever enabled. For more details about this,
the ratio of the overlap to the minimum fragment size exceeds that please refer to the "About Setting Global
Preferences" section in the Rocky User
value, the particle will break, regardless of the energy values on the Manual.
particle at that moment. The minimum fragment size is a user input
listed as Minimum Size on the Breakage / Fragment Distribution sub-tab
of the particle being defined.

© 2020, esss - all rights reserved


dem technical manual 27
ROCKY

When a particle breaks because of a large overlap, the resulting


fragment size distribution is determined according to the same
model chosen for the primary breakage model, Gaudin-Schumann or
incomplete beta function. The value of the t10 parameter is computed
with the respective value of contact energy at the moment of breakage.

2.4.4 Fragment size distribution models

Regardless of the type of instantaneous fragmentation model that is


selected, once the particle reaches the breakage point, Rocky generates
fragments preserving both the total particle mass and volume, using
one of the two available distribution models: Gaudin-Schumann or
incomplete beta function.
The number and size of fragments will be dependent upon the
particle’s original size, on the value of the t10 parameter, and the value
of the minimum admissible fragment size. This last parameter is
controlled by two user inputs, listed in the Rocky UI as Minimum Size
and Minimum Size Ratio , respectively. The first one allows users to
specify the minimum size in absolute terms, whereas the second
does it in relative terms. The minimum size allowed for fragments
in Rocky will be the larger value between the Minimum Size value
and the product of the Minimum Size Ratio and the size of the particle
being broken. When generating the fragments at a breakage event,
no fragment is allowed to have a size smaller than the resulting value
from that operation.
Allowing the generation of too small fragments would be inad-
visable mainly because the computation time may increase beyond
reasonable limits. The reason for this is twofold: the total number of
objects that must be tracked in the simulation may grow exponentially,
whereas the timestep may need to be reduced significantly in order
to keep the simulation process stable.40 A side effect of limiting 40
Roughly speaking, the stable timestep
is proportional to the size of the smaller
the minimum size of fragments is that the actual size distribution particle in a simulation.
satisfies the prescribed size distribution41 only for fragments larger 41
Given by equation (2.54) for the
to approximately 2.5 times the specified minimum size. Gaudin Schumann distribution and by
equation (2.55) for the incomplete beta
In the same line of preventing the generation of too small distribution.
fragments, an additional safeguard parameter is defined as the
Minimum Volume Fraction for Fragment Disabling . When a generated frag-
ment is detected to have an actual volume smaller than the value
of that parameter multiplied by the minimum volume allowed for a 42
This minimum volume is the volume
fragment,42 such fragment does not enter into the simulation. of an equivalent sphere of diameter
equal to the minimum size defined
above.

© 2020, esss - all rights reserved


dem technical manual 28
ROCKY

2.4.4.1 Gaudin-Schumann

The full fragment size distribution is determined from the value of t10
by assuming a Gaudin-Schumann distribution with unitary slope:
x
Y = 10 t10 (2.54)
L
where:

• Y is the cumulative percentage of the mass of passing fragments.


• x is the screening sieve size.
• L is the size of the broken particle.

2.4.4.2 Incomplete beta function

The full fragment size distribution is determined from the value of


t10 by assuming an incomplete beta function distribution:43 43
Barrios, G. K. P., Carvalho, R. M.,
and Tavares, L. M. (2011). Modeling
t10 breakage of monodispersed particles in
100
Z
100
tn (t10 ) = R 1 x αn −1 (1 − x ) β n −1 dx (2.55) unconfined beds. Minerals Engineering,

0 x αn −1 (1 − x ) β n −1 dx 0 24:308–318

where:
• tn is the percentage of fragments passing a screen size of 1/nth of
the original size L.
• αn and β n are model parameters fitted to experimental data.
In Rocky, the user can introduce values for these parameters,
corresponding to different values of n, in a table provided in the
corresponding data editor panel.

1.2 0.19 7.78

1.5 0.56 7.51

2 0.78 5.55

4 1.12 3.01

25 1.17 0.54
5%
10% 50 1.43 0.40

20% 75 1.92 0.42


40%

Figure 2.5: Examples of fragment size


distributions for different values of t10 .

Figure 2.5 shows examples of fragments size distributions obtained


44
Carvalho, R. M. and Tavares, L. M.
(2013). Predicting the effect of operating
with parameters fitted to experimental data for limestone.44 Marker and design variables on breakage rates
using the mechanistic ball mill model.
Minerals and Engineering, 43:91–101

© 2020, esss - all rights reserved


dem technical manual 29
ROCKY

points in the plot correspond to values computed with equation (2.55),


using the values of αn and β n reproduced in the same figure. In Rocky,
intermediate values of the fragment size distribution are determined
by means of linear interpolation of the values computed with equation
(2.55).

2.5 Wear model


Rocky can be used also as a tool for the prediction of abrasive wear
of solid surfaces due to the action of impacting particles. The wear
model implemented in Rocky is based on Archard’s wear law45 . This 45
Archard, J. F. (1980). Wear theory and
mechanisms. In Wear control handbook.
phenomenological law relates the volume loss of material to the work American Society of Mechanical Engi-
done by the frictional forces on the surface of the material. Archard’s neers

law is usually expressed as:46 46


Qiu, X., Potapov, A., Song, M., and
Nordell, L. (2001). Prediction of wear
Fτ sτ of mill lifters using discrete element
V=k (2.56) method. In 2001 SAG Conference Pro-
H
ceedings
where:

• V is the total volume of material worn from the surface.


• Fτ is the tangential force exerted on the surface.
• sτ is the sliding distance on the surface.
• H is the hardness of the material subjected to wear.
• k is a dimensionless empirical constant.

For the purpose of implementation in Rocky, Archard’s law is


considered in the incremental form:

∆V = C ∆Wτ (2.57)
where:

• ∆V is the volume of material worn during a simulation timestep.


• ∆Wτ is the tangential or shear work done by the particles colliding
with a surface during the same timestep.
• C = k/H is a constant provided by the user. In Rocky, every
imported boundary in a problem can be defined with a different
value of the constant C, listed in the UI as Volume/Shear Work Ratio .

Figure 2.6 illustrates schematically the application of equation (2.57)


in Rocky. Since all boundaries in Rocky are defined as triangulated
surfaces, the effect of removing the volume ∆V is achieved displacing
their vertices inwards. The distance that every vertex is displaced is
calculated in order to make the volume variation equal to the value

© 2020, esss - all rights reserved


dem technical manual 30
ROCKY

Particles doing shear work


Volume worn from the surface

Boundary

(a) Time (b) Time

Figure 2.6: Schematic depiction of the


wear of a surface.
of ∆V calculated with equation (2.57). It is worth noting that during
this process the topology of the boundary triangulation is maintained
unchanged. Because of this, in order to capture adequately geometry
changes, it is desirable that regions expected to suffer wear present
47
Users must be aware, however, that
the more refined the boundary triangu-
triangulations sufficiently refined.47 lation, the longer will be the processing
time.
The tangential work done by the particles colliding with the
boundary is computed triangle by triangle. Vertices adjacent to
triangles that receive more shear work will displace more and,
consequently, these regions will show more wear in a simulation. The
tangential work per triangle is calculated by adding the tangential
work contributions of all collisions occurring in a triangle during a
simulation timestep.
In general, the actual period of life of wear in industrial processes
is much longer than what is usually reasonable to simulate using
DEM. This problem can be circumvented by shortening the simulation 48
Qiu, X., Potapov, A., Song, M., and
time while increasing the wear rate C several orders of magnitude. A Nordell, L. (2001). Prediction of wear
of mill lifters using discrete element
suitable value of C can be determined by following two rules:48 method. In 2001 SAG Conference Pro-
ceedings
1. C must satisfy the constraint of proportionality of equation
(2.57). In other words, increasing C by a factor and reducing
the simulation time by the same factor must preserve the wear
characteristics.

2. The entire simulation time must be sufficiently long, so the


statistical variance associated with discrete events is small in the
DEM output data.

Both rules are violated when extremely large values of C are used,
producing fast changes in geometry that strongly disturb the motion
of particles.

© 2020, esss - all rights reserved


dem technical manual 31
ROCKY

2.6 Thermal model


If the thermal model is enabled in Rocky, an additional equation for
thermal energy balance is solved along with the equations governing
the motion of particles. In the current implementation of Rocky, the
temperature is assumed to be uniform in standard single-element
particles (non-flexible). This means that no radial nor circumferential
temperature variation is admitted inside any of these particles, which
is reasonable if they are small and/or made of highly conductive
materials.49 On the other hand, for multi-element flexible particles, 49
Vargas, W. L. and McCarthy, J. J. (2001).
Heat conduction in granular materials.
which are formed by connected elements as described in chapter AIChE Journal, 47(5):1052–1059
3, the temperature may have variation between elements. However,
the temperature of each element making up a flexible particle is
considered to be uniform also.
In Rocky, the temperature variation of a particle or element e over 50
In this and the subsequent equations,
time is determined by solving the differential equation:50 e stands for either an entire particle or
a single element in a multi-element flex-
dTe ible particle, depending upon whether
me ce = q̇e (2.58) the particle being considered is a stan-
dt
dard (rigid) particle or a flexible one.
in which, me and ce are, respectively, the mass and the specific heat
of the particle or element e, while q̇e is the overall heat transfer rate
between it and its surroundings. For instance, in all DEM problems,
q̇e accounts for the conduction heat transfer that occurs during the
contact with other particles or walls. Additionally, if the problem
being solved includes fluid flow, the heat transfer between the particle
and the fluid phase, q̇ f →e , is included in q̇e as well.51 Moreover, if e 51
For a description of the correlations
considered in Rocky to model q̇ f →e ,
is an element of a flexible particle, the heat transfer across the inter- please refer to the DEM-CFD Coupling
element joints must be considered also. Then, for the most general Technical Manual.

case we can write:

Nc, e Nj, e
q̇e = ∑ q̇c→e + ∑ q̇ j→e + q̇ f →e (2.59)
c =1 j =1

where q̇c→e is the instantaneous heat transfer rate from another


particle or a wall through a contact c, whereas q̇ j→e is the instan-
taneous heat transfer rate from the other element of the same particle
connected to e through a joint j. In this equation, Nc, e is the number
of contacts that e has with other particles or walls at a given time,
while Nj, e is the number of joints at which e connects itself with other
elements.52 52
If e is not a single element of a flexible
particle, the second term on the right-
hand side of equation (2.59) is ignored
completely.

© 2020, esss - all rights reserved


dem technical manual 32
ROCKY

2.6.1 Contact heat transfer rate

In Rocky, we consider that the heat conduction between a particle in


contact with another particle or a wall occurs exclusively through the
contact area between them.53 Considering that the temperature of a 53
The same mechanism of heat transfer
described in this section occurs also
given particle is Te and the temperature of the contacting particle or when a single element of a flexible parti-
wall is Tη , the heat transfer rate q̇c→e is modeled through the linear cle collides with another non-connected
element, or a single-element (rigid) par-
expression: ticle, or a wall. The heat conduction be-
q̇c→e = Hc ( Tη − Te ) (2.60) tween two elements connected through
a joint is explained on the next section.

where Hc is the contact conductance. Batchelor & O’Brien54 proposed 54


Batchelor, G. K. and O’Brien, R. W.
(1977). Thermal or electrical conduction
the following expression for the thermal contact conductance between through a granular material. Proc. R. Soc.
two particles of uniform temperature, based on the analytical solution Lond. A., 355:313–333

of an analogous irrotational flow problem:

Hc = 2 k c a (2.61)

where k c is the equivalent thermal conductivity of the contact and a


is the radius of the contact surface, which is depicted schematically
in Figure 2.7. In order to determine the value of the contact radius,
Batchelor & O’Brien considered the Hertz theory,55 according to 55
Hertz, H. (1882). Über die berührung
fester elastischer körper (On the contact
which we have: of elastic solids). J. Reine Angewandte
1
3 Fn R∗

3 Mathematik, 92:156–171. English transla-
a= (2.62)
4 E∗ tion, Macmillan, London, 1896

where:

• Fn is the instantaneous normal contact force between the two


particles or the particle and the wall.
• R∗ is the equivalent radius, defined in equation (2.15).
• E∗ is the effective Young’s modulus, defined in equation (2.14).

Figure 2.7: Heat transfer between two


colliding particles.

Equations (2.61) and (2.62) are used in Rocky at particle-boundary


contacts as well, considering the appropriate definitions for the
involved parameters. Regarding the equivalent thermal conductivity

© 2020, esss - all rights reserved


dem technical manual 33
ROCKY

used in equation (2.61), as the thermal conductivities of the particle


e and its neighbor can be different in a general case, in Rocky it is
defined as the harmonic mean:

2
kc = 1 1
(2.63)
ke + kη

where k e and k η are the thermal conductivities of the two contacting


entities.

2.6.2 Joint heat transfer rate

The heat transfer rate between two elements of a flexible particle


connected by a virtual joint is modeled in Rocky as:

q̇ j→e = Hj ( Tη − Te ) (2.64)

where Te and Tη are the temperatures of the two connected elements,


while Hj is the joint equivalent conductance, given by:

1 1 1
= + (2.65)
Hj He Hη

where H p and Hη are the partial thermal conductances at each side


of the joint, approximated respectively as:

`e
He = (2.66)
k j Aj

Hη = (2.67)
k j Aj

In these expressions, `e and `η are the distances from the centroid


of the joint to the centroid of the respective element, as shown in
Figure 2.8, in a flexible fiber example. Additionally, k j is the thermal
conductivity of the joint, defined as:

j
k j = rk k (2.68)

j
where k is the thermal conductivity of the particle’s material and rk
is a user input parameter, listed in the Rocky UI on the Composition
sub-tab for the Particle set as Conductivity Ratio . This parameter allows
users to make fine adjustments to the joint’s thermal conductivity
without modifying the particle’s thermal conductivity, which is used
in the calculation of the contact conductance as well. Allowed values
for the conductivity ratio are in the range from 0.01 to 100.

© 2020, esss - all rights reserved


dem technical manual 34
ROCKY

Figure 2.8: Heat transfer between two


elements in a flexible fiber.
In equations (2.66) and (2.67), A j is the effective area of the joint.
In most cases, A j will be the area of the section that connects two
elements. For straight fibers, it will be the area of the circular section
of the cylindrical part of the elements. For shells, it will be the area
of a rectangular face connecting two prismatic elements.56 For solid 56
See Figure 3.6 for an example of a
connection face in a flexible shell.
particles, it will be the area of the triangular face connecting two
tetrahedral elements.57 57
Figure 3.7 depicts an illustrative exam-
ple of a connection face in a solid flexible
The case of multi-branched custom fibers is the one that poses particle.
certain difficulties for defining A j . The main difficulty is that several
sphero-cylinder elements can meet at a given point, each one having
a different diameter, as depicted in the example of Figure 2.9. As an
element may be connected to more than one element at one of its
ends, this reduces the effective area for transferring heat to any of the
connected elements at that place. Therefore, we define the following
effective area for those cases:

Amin
Aj = j
(2.69)
Ne − 1

where Amin is the smallest cross sectional area between the two
j
elements connected by the given joint, while Ne is the number of fiber
elements that meet at the same point were that joint is located.

Figure 2.9: Example of a multi-branched


custom fiber. In this case, the number
j
of coincident elements is Ne = 4. For
the joint between the green and blue
elements, Amin is the cross sectional area
of the blue element (the one with the
smallest diameter).

© 2020, esss - all rights reserved


dem technical manual 35
ROCKY

2.6.3 Thermal conduction correction models

Decreasing stiffness in order to speed up a simulation58 has the 58


Please refer to section 5.2.4 for a
complete description of this technique.
undesired side effect of abnormally increasing the predicted heat
exchange. When the contact stiffness is decreased using the
Numerical Softening Factor , the contact area and the collision time are
increased, both causing an artificial rise of the heat transferred
through a contact. In order to counteract those adverse effects, Rocky
includes two correction strategies proposed initially by Morris et al.59 . 59
Morris, A. B., Pannala, S., Ma, Z., and
Hrenya, C. M. (2016). Development
The Morris et al. Area correction is valid for static systems. When of soft-sphere contact models for ther-
enabled by the user in the Thermal tab of the Physics panel, this mal heat conduction in granular flows.
AIChE Journal, 62(12):4526–4535
correction model replaces the reduced Young’s modulus of particles
or boundaries by the actual values specified by the user60 for the 60
The actual Young’s modulus values
are defined by the user in the Materials
respective materials, when calculating the contact radius using panel.
equation (2.62).
The Morris et al. Area+Time correction is indicated for simulations
with non-resting particles. This model extends Morris et al. Area by also
applying a further correction factor α to the area-corrected contact
radius during collisions in order to mitigate the adverse thermal effect
caused by artificially longer contact times between softened particles.
This model was deduced by considering the theoretical exchanged
heat rate during a collision between two particles for the actual and
softened cases, and provides the following correction factor for the
collision time:
 2
tc,actual 3
α= (2.70)
tc,sim
where:

• tc,sim is the collision time between the softened particles.


• tc,actual is the collision time between the actual particles.

The correction factor α also equals to the ratio of the actual and
area-corrected thermal contact conductances between the colliding
particles. Therefore for a Hertzian collision model it can also be
interpreted as a multiplier to the instantaneous, area-corrected contact
radius between softened particles during the simulation61 , 62 . 61
This interpretation is accurate for elas-
tic binary collision systems as would be
Still considering a Hertzian collision model, equation (2.70) can be expected in dilute granular flows.
simplified to: 62
For non-Hertzian collision models, the
4
α = (NSF) 15 (2.71) errors in the predicted heat transfer
resulting from this assumption are quite
small, as demonstrated by Morris et. al
where NSF is the Numerical Softening Factor of the simulation. Equation for the linear spring dashpot collision
(2.71) is employed by Rocky when the Morris et al. Area+Time correction model.

model is enabled by the user.

© 2020, esss - all rights reserved


dem technical manual 36
ROCKY

2.7 Coarse-Grain Model


Coarse-Grain Modeling (CGM) is a method used to reduce the total
number of particles in a simulation by using larger particles to
represent groups of the original (smaller) particles. Modifications on
the contact, adhesion and other models are necessary to compute
the interactions on this scaled-up system in a way that will allow
the dynamics of this scaled-up system to be similar with the original
system.
The scaled-up particle, usually referred to as a parcel, is defined
by the original particle characteristics (material, shape, size, etc.) and
by a scale-factor. Because these characteristics can be unique per
particle group, each particle group can have its own scale factor. This
CGM Scale-Factor f CGM multiplies the original particle size to define
the parcel size and, consequently, the number of original (smaller)
particles the parcel represents. Due to the scale-factor to volume
3
relationship, the parcel represents f CGM original particles. Figure 2.10
shows the concept of the parcel.

Figure 2.10: A collection of small (orig-


inal) particles is replaced by a larger
CGM parcel (scaled-up particle). The parcel
mass is the sum of the original particle
scaling masses. In this example f CGM = 2.
fCGM L

In case a particle size distribution is used for the original particle


size of the group, the same scale factor will be used for all particles
generated on that group. The final PSD will have the same shape
of the original particle size distribution, but each parcel in the DEM
simulation will have the size scaled by f CGM times the original particle
size.
It is important to note that each parcel will represent a collection
of smaller particles with the same characteristics, including size. So,
in cases where a PSD is defined, each parcel will represent collection
of original particles with a specific size, not a group of particles with
the entire original size distribution.
Besides reducing the number of particles used within the simula-

© 2020, esss - all rights reserved


dem technical manual 37
ROCKY

tion, the CGM approach used in Rocky also allows the usage of larger
time steps for the calculations. The increase of the time step comes
from the scaling-up of the overlap between two particles in contact.
This overlap is scaled up by the factor f CGM . As the velocities of the
parcels are similar to the velocities of the original particles, the contact
time is also increased by a factor of f CGM .
The Coarse-Grain Model approach used by Rocky is based on the
work of Bierwisch et al.63 . The original work defines a procedure to 63
Bierwisch, C., Kraft, T., Riedel, H., and
Moseler, M. (2009). Three-dimensional
identify modifications that need to be applied to the contact force discrete element models for the granular
and adhesion models to be used on the scaled-up system. Using statics and dynamics of powders in
cavity filling. Journal of the Mechanics
the modified versions of these models, when two parcels collide, the and Physics of Solids, 6257:10–31
3
kinetic energy variation due to the contact will be the same as f CGM
contacts of the original particles 64 . 64
The velocities of the parcels after
contacts are expected to be similar to
The kinetic energy variation of the system ∆Ek , is defined according those on the original system but other
to Equation (2.72). quantities, such as forces, stresses, etc.,
will be scaled up on the resuts. The scal-
ing up of these quantities must be taken
m∗  2 
into account when post-processing sim-
∆Ek = ṡb − ṡ2a (2.72)
2 ulation results using CGM.

where, ṡb and ṡ a are the relative velocities of the particles before and
after the contact, respectively, and m∗ is the effective mass of the
contact.
Considering a constant effective contact mass, in order to conserve
the kinetic energy variation, the velocity variation of the scaled-up
system during the contact must be kept the same as in the original
system. Figure 2.11 illustrates this concept. On the left side of the
image, the original system is shown. On this system, a number of
contacts of the original particles are happening. The velocities before
and after the contact are shown. The two groups of particles are scaled-
up, each group of particles (blue and green) will be represented by a
single parcel on the scaled-up system. On the right side of the image,
this scaled-up system is represented. The velocities of the scaled-up
system are supposed to be the average velocity of the particles of the
original system, before and after the contact.
As the mass of the parcels is equal to the sum of the mass of the
original particles, if they have the same average relative velocity of
the original system, then the kinetic energy is conserved.
The procedure defined within the work of Bierwisch et al. uses
dimensional analysis on the momentum-impulse equation, given by
(2.73) for the contact normal direction.

m∗ ∆ṡn = ( Fn + Fn,adh ) ∆t (2.73)

© 2020, esss - all rights reserved


dem technical manual 38
ROCKY

pre-contact Figure 2.11: Eight contacts of the origi-


nal system are represented by a single
pre-contact contact on the scaled-up system with
f CGM = 2.

averaged velocities
before contact
CGM
scaling
post-contact

post-contact

averaged velocities
after contact

The modifications necessary on each contact force and adhesion


model for the scaled system are defined by the dimensional analysis
when substituting the normal force, Fn by the desired contact normal
force model expressions and Fn,adh by equations of the chosen
adhesive model. This process is repeated for each contact force
and adhesive model available. Typically the modifications needed are
just the inclusion of the scale-factor multiplying some of the physical
properties used on the calculations. The same procedure can be used
on the tangential contact force models.

2.7.1 Contact frictional forces

When performing the dimensional analysis to determine the contact


frictional force model modifications necessary for the scaled system, it
was discovered that these contact force models implemented in Rocky
are self-similar, which means that they do not need to be modified to
satisfy the Bierwisch et al. model constraint. These models are:

• Normal Contact Force:

– Hysteretic Linear Spring

© 2020, esss - all rights reserved


dem technical manual 39
ROCKY

– Linear Spring Dashpot

– Hertzian Spring Dashpot

• Tangential Contact Force:

– Linear Spring Coulomb Limit

– Coulomb Limit

– Mindlin-Deresiewicz

2.7.2 Adhesive forces

The Constant adhesive force model needed to be adjusted according


to the dimensional analysis. The CGM version of this model calculates
the adhesive force as:


 0 if − sn > δadh
Fn,adh = (2.74)
g min(m1 , m2 ) if − sn < δadh
 f
adh f CGM

The Linear adhesive force model does not require any modification.
All other adhesive force models, including Leeds , JKR and
Velocity Dependent , are not currently compatible with CGM.

2.7.3 Contact heat transfer

For thermal enabled simulations, the model constraint is that the


temperature variation experienced by two parcels exchanging heat
during contact should be the same as the temperature variation of
the one that would happen during a collision between two original
particles the parcels represent.
Figure 2.12 shows this constraint. Again, on the left side of
the image the original system is shown. A number of contacts of
original particles are happening (just one contact is shown, the ellipsis
indicates that there are a number of this same contact happening).
One set of particles have a higher temperature (indicated by the red
color) and the other set have lower temperatures (indicated by the
blue color). During the contact, the particles exchange heat and after
the contact they have closer temperatures (indicated by the lighter
colors of each group). On the scaled-up system, the parcels initially
have their temperature defined as the averaged temperature of the
collection of original particles they represent. The thermal constraint

© 2020, esss - all rights reserved


dem technical manual 40
ROCKY

for CGM is that their temperatures at the end of the contact, should
be the same as the post-contact averaged temperature of the collection
of the original particles.

original scaled-up Figure 2.12: Eight contacts of the orig-


system system inal system with energy transfer are
represented by a single contact on the
scaled-up system with f CGM = 2.
pre-contact

pre-contact

averaged temperatures
... before contact

post-contact post-contact

...

averaged temperatures
after contact

In order to identify the changes necessary on the contact heat


transfer models, the same procedure of dimensional analysis was
applied to the particle thermal energy conservation equation, given
by equation (2.75).

mc∆T = q̇ (2.75)

Here, c is the specific heat of the particle material. The heat transfer
rate q̇ is calculated applying the contact heat conduction model.
The dimensional analysis proved that the thermal conduction
model implemented in Rocky is also self-similar, therefore, no
modification is needed in order to make it consistent with the CGM
approach.

© 2020, esss - all rights reserved


dem technical manual 41
ROCKY

2.7.4 Radl et al. kinetic energy dissipation model

When particles are moving rapidly (not on a bed or settled over a


conveyor belt) with fast contacts, the usage of Coarse Grain Modeling
can underestimate the total energy dissipation on the system because
of missed contacts on the scaled-up system. Figure 2.13 illustrates
these missed contacts.
The first type of missed contacts is the intra-parcel contact. The
parcel velocity is the the average velocity of the original particles it
represents, but each individual particle could have a different velocity
on the original system. These different velocities lead to contacts
between particles on the original system. These contacts are shown in
the original system, marked by smaller circles inside each group.
The second type of contact missed by the scaling up are some
inter-parcel contacts. When two groups of particles pass by each other
closely, some particles on the periphery of each group may collide,
as highlighted by the smaller circle between the two groups on the
original system. Figure 2.13 shows that in this case, the contacts were
missed on the scaled-up system.

Figure 2.13: Two groups of particles


scaled up with CGM. Particles within
the smaller circles represent four con-
tacts that were missed due to the scaling-
up process: three due to intra-particle
collisions and one due to an inter-parcel
collision. f CGM = 2.

CGM
scaling

To better represent the outcome of these missed collisions, the


Radl et al.65 model introduces a correction to the parcels’ velocity 65
Radl, S., Radeke, C., Khinast, J. G.,
and Sundaresan, S. (2011). Parcel-based
at each time step. This velocity correction simulates the effect of the approach for the simulation of gas-
energy dissipated by the missed collisions. The corrected velocity, v0 particle flows. In Proceedings of the
8th international Conference on CFD in
the Oil & Gas, Metallurgical and Process
Industries., pages 124–134

© 2020, esss - all rights reserved


dem technical manual 42
ROCKY

is calculated as shown in equation (2.76).

∆t
v+ 2τD v̄
v0 = ∆t
(2.76)
1 + 2τD

In equation (2.76) τD is the particle relaxation time, or damping


time, and v̄ is the mass weighted averaged velocity of the parcels in
the vicinity.
The particle relaxation time is calculated using equation (2.77):

 √
1 1 2 αp
=h 1 − g0
τD f CGM 3π L332
4
(2.77)

∑ p ητD (1 − ητD ) N
L+ L32
2 (v − v̄)2
 2
L+ L32
∑p N 2 (v − v̄)

In this equation, α p is the volume fraction of particles within a


given search radius, L32 is the Sauter mean particle size, and N is the
number of particles each parcel represents.
Also in equation (2.77), h is the shutoff function, designed to turn
off the model when the particles are on a settled bed of particles, or
moving on a bed with slow contacts, given by equation (2.78). g0 is
the radial distribution function, which provides the probability of the
missed contacts. This function is shown in equation (2.79). Finally,
ητD is the damping function, which represents the fraction of energy
lost on a missed contact, given by equation (2.80).
 !8 
α p,cp − α p
h = min 1,  (2.78)
α p,cp − α p,off

α p,cp
g0 = (2.79)
α p,cp − α p

1+ε
ητD = (2.80)
2
In equations (2.78) and (2.79), α p,cp is the Close Packing Volume Fraction
of the particles, the maximum volume fraction the particles are
allowed to reach, which is defined by the user. In equation (2.78),
α p,off is the particle volume fraction value at which the model will be
shut off, which is automatically calculated as 95% of the close packing
volume fraction.
The calculation of the averaged quantities and local volume fraction
are based on the properties of near parcels. These near parcels are

© 2020, esss - all rights reserved


dem technical manual 43
ROCKY

identified using a search algorithm similar to the one used to identify


particle neighbors for contact calculations, described in section 5.1.
This near parcel search uses a spherical region around each parcel,
with the radius defined by the radius of the sphere that circumscribes
the largest parcel and a Search Distance Multiplier , which is provided by
the user. The parcels that are inside the sphere centered on a given
parcel are considered neighbors within the calculations. This list of
near parcels is updated when the parcels travel a distance larger than
a limit value, this value is calculated using the same circumscribed
sphere and an Update Distance Multiplier , which is also provided by the
user.

© 2020, esss - all rights reserved


dem technical manual 44
ROCKY

3 Particle types in Rocky


As of v4.1, Rocky offers three different particle shape categories
–Fiber, Shell, and Solid– which are described in detail below.

3.1 Fiber particle shapes


Fiber is a category of particle shapes that are mainly one-dimensional,
i.e., the main geometry of a fiber can be described by a line in a
three-dimensional space. Rocky comes with a Straight Fiber by default,
but Custom Fiber shapes can be defined by the user in a .txt file and
then imported into Rocky.
By default, single-element Fiber shapes are treated as solid inflexible
objects. However, Fiber particles can be treated as flexible if they
are defined as composed by multiple sphero-cylinder elements, as
described below.

3.1.1 Flexible straight fibers

A flexible fiber is generated in Rocky when the option Multiple Elements


is selected in the Composition sub-tab associated with a straight fiber
particle shape. The modeling of flexible fibers in Rocky follows the
approach described by Guo et al.1 . In this approach, a flexible fiber 1
Guo, Y., C., W., Curtis, J. S., and Xu,
D. (2018). A bonded sphero-cylinder
is built by connecting sphero-cylinders by means of joint entities, as model for the discrete element simula-
depicted schematically in Figure 3.1(a). A joint is an auxiliary entity tion of elastic-plastic fibers. Chemical
Engineering Science, 175:118–129
with elastic and viscous properties that connects two adjacent sphero-
cylinder elements. At the beginning of a simulation, the centers of the
hemispherical ends of adjacent elements coincide and the joints are
undeformed. However, when the a fiber interact with other fibers or
boundaries, the relative movement between the elements may produce
linear and angular deformations on the joints.2 In response, forces 2
In this approach, only the joints can
deform, while the sphero-cylinder el-
and moments are induced and exerted on the adjacent elements to ements are considered perfectly rigid,
resist the normal, tangential, bending, and torsional deformations, as at least from the point of view of the
interaction between them.
can be seen schematically in Figure 3.1(b).

© 2020, esss - all rights reserved


dem technical manual 45
ROCKY

Sphero-cylinder element

(a)

(b)

Figure 3.1: (a) Schematic representation


of a multi-element flexible fiber. (b)
Each force and moment is modeled in Rocky as the sum of two Forces and moments that a joint exerts
on a fiber element.
parts: an elastic one and a damping one. For instance, the normal
and shear forces are calculated, respectively, according the following
expressions:

j j j
Fn = Kn srel rel
n + Cn vn (3.1)
j j j
Fτ = Kτ srel rel
τ + Cτ vτ (3.2) Equation (3.2) is written as a vector
equation because, in general, srel τ and
vrel
τ will not be parallel, although both
where:
are always perpendicular to the normal
j j direction. This is clarified below.
• Kn and Kτ are the normal and tangential stiffnesses associated to a
joint, respectively.
j j
• Cn and Cτ are the normal and tangential damping coefficients
associated to a joint, respectively.

• srel rel
n and sτ are, respectively, the normal and tangential relative
displacements between the center points of the hemispherical ends
of the two elements connected by a joint (see Figure 3.3).

• vrel rel
n and vτ are, respectively, the normal and tangential relative
velocities between the center points of the hemispherical ends of
the two elements connected by a joint. The absolute velocity at
each center point is the sum of the velocity of the element’s center
of mass and the tangential velocity due the rotation of the element
around that center of mass.

Figure 3.2 illustrates the determination of the normal direction


for the vector decomposition in the above equations. If â1 and
â2 are unit vectors parallel to the direction of the two elements
connected by a joint, the normal direction is defined by the unit
â1 +â2
vector n̂ = |â1 +â2 |
. As shown in Figure 3.3, the relative displacement
vector srel is then decomposed into two perpendicular components,

© 2020, esss - all rights reserved


dem technical manual 46
ROCKY

Figure 3.2: Relative displacement vector


and determination of the normal direc-
tion.

the normal component:

srel rel
n = s · n̂ (3.3)
n
and a tangential component:
n
srel
τ =s rel
− srel
n n̂ (3.4)

The relative velocity vector is decomposed in a similar way. In


general, however, srel rel
τ and vτ will not be parallel, although both Figure 3.3: Decomposition of the relative
always will be perpendicular to n̂. displacement vector.

The normal and tangential stiffnesses for a joint are defined,


respectively, as:
j Ej A
Kn = (3.5)
`j
j
j Kn
Kτ = (3.6)
2(1 + νj )

where:

• Ej is the Young’s modulus of the joint.


• A is the cross sectional area of the joint, considered numerically
equal to the cross sectional area of the cylindrical part of the fiber
elements.
• ` j is the length of the joint. In general, this length is considered
equal to the average length of the sphero-cylinders that form the
fiber.
• νj is the Poisson’s ratio of the joint, which is considered equal to
the Poisson’s ratio of the material associated to the fiber.

The Young’s modulus of the joint is related to the Young’s modulus


of the material associated to the fiber, E p , by means of the simple
expression:

© 2020, esss - all rights reserved


dem technical manual 47
ROCKY

j
Ej = r E E p (3.7)
j
where r E is a user input parameter in Rocky, listed in the UI as
Young’s Modulus Ratio . This parameter allows the user to fine-tune the
flexibility of joints. Allowed values are in the range from 0.01 to 100.
The normal and tangential damping coefficients for the joint are
determined, respectively, by the expressions:
q
j j
Cn = 2η j m ∗ Kn (3.8)
q
j j
Cτ = 2η j m∗ Kτ (3.9)

where:

• m∗ is the effective mass of the joint, which is determined with


equation (2.10), considering the masses of the two adjacent sphero-
cylinder elements.

• η j is a user input parameter listed as Damping Ratio in the Rocky UI.


A value on the 0 < η j < 1 interval means that the system will be
underdamped, therefore, decaying oscillations may be observed in
the relative displacement between elements. The higher the value
of η j , the faster will be the decay of oscillations.

The bending and torsional moments acting at the ends of the


joint are calculated following a similar approach. The expressions
considered are, respectively:

j j
j Kn I C I rel
Mben = θ + n ωben (3.10)
A ben A
j j
j Kτ J C J rel
Mtor = θtor + τ ωtor (3.11)
A A

where:

• I and J are the area moment of inertia and the polar area moment
of inertia, respectively, of the circular cross section of the joints.

• θben and θtor are, respectively, the bending and torsional angular
displacements between the two adjacent sphero-cylinder elements.
rel and ω rel are, respectively, the bending and torsional relative
• ωben tor
angular velocities between the two adjacent sphero-cylinder
elements.

© 2020, esss - all rights reserved


dem technical manual 48
ROCKY

3.1.2 Flexible custom fibers

With custom fibers, sphero-cylinder segments can be arranged arbi-


trarily forming complex multi-branched geometries. The geometry
of custom fibers must be specified by means of a TXT file, in which
every row defines a segment of the fiber geometry.3 As shown in 3
See also the "Import Custom Particle
Shapes" topic in the Rocky User Manual.
the example depicted in Figure 3.4, the first 6 numbers in every row
correspond to the Cartesian coordinates of the two end points of a
segment. The seventh number defines the segment diameter. The
optional eighth number is a multiplier of the Young’s modulus for
the segment. If it is not explicitly defined for a segment, its value is
considered equal to 1. This parameter allows to define a custom fiber
with nonuniform flexibility, as described below.

0.1
Figure 3.4: Example of the specification
of a custom fiber with nonuniform
flexibility.
0.4 x1 y1 z1 x2 y2 z2 d M

0.0 0.0 0.0 0.1 0.0 0.0 0.02


1 1
0.1 0.0 0.0 0.2 0.0 0.0 0.02

0.1 0.0 0.0 0.15 0.1 0.0 0.02 0.1

0.1 0.0 0.0 0.15 -0.1 0.0 0.02 0.3

0.1 0.0 0.0 0.15 0.0 0.1 0.02 0.5


0.3
y 0.1 0.0 0.0 0.15 0.0 -0.1 0.02 0.4

x
0.5
z

The modeling of flexible custom fibers is exactly the same described


in the previous section for flexible straight fibers. Between any pair of
connected segments, an elastic-viscous joint is considered, that exerts
forces and moments on the segments that oppose to translational
and angular deformations. The major difference is that, as stated
before, custom fibers can be defined with nonuniform flexibility. This
effect can be be achieved by specifying different values to the Young’s
modulus multipliers for the segments that form the fiber.
The flexibility of a joint is determined mostly by the value of the
Young’s modulus attributed to it. In Rocky, this property is computed
as:
j
Eb = Mi,j r E E p (3.12)
where:

• E p is the Young’s modulus of the particle’s material.


j
• r E is a global multiplier for the whole custom fiber, which is

© 2020, esss - all rights reserved


dem technical manual 49
ROCKY

listed in the Rocky UI as Young’s Modulus Ratio and is located on the Mi


Composition sub-tab of the corresponding particle group.
Mj
• Mi,j is the harmonic mean of the multipliers specified for the two
segments associated to the joint:

2Mi M j
Mi,j = (3.13)
Mi + M j Figure 3.5: Joint in a custom fiber.

where, as depicted schematically in Figure 3.5, Mi and M j are the


multipliers specified for the two adjacent segments, respectively.
If the segments of a custom fiber are further subdivided into
subsegments,4 the value of Mi,j for the joints between those 4
This will happen if the parameter
is specified with
Target Number of Elements
subsegments will be equal to the value of the multiplier defined
a high value in the Composition sub-tab
for the segment that contains them. for the particle group.

3.2 Shell particle shapes


Shell particles can be thought of as hollow open containers of uniform
thickness, whose shape is determined by a two-dimensional surface.
No default Shell particles come with Rocky, but custom Shell shapes
can be defined from an open tessellated surface imported as a STL
file.

3.2.1 Flexible shells

By default, Shell particles are treated in Rocky as inflexible objects.


However, if the option Multiple Elements is selected in the Composition
sub-tab, flexible Shell particles will be generated. The modeling of
flexible Shell particles in Rocky is an extension of the model of flexible
fibers described previously in section 3.1.1.
A flexible Shell is formed by thin prismatic elements. Each one of
those elements arises from a triangle in the open surface supplied to
Rocky as an STL file, for defining the particle shape. The thickness
of those prismatic elements is the thickness defined by the user for
the whole Shell particle. The prismatic elements are connected among
them by rectangular faces, as depicted schematically in Figure 3.6.
In the undeformed state of a Shell particle, the corresponding
faces of two adjacent elements will coincide if the shell is locally
flat. If the Shell has curvature, at least the major center line of two
corresponding faces will coincide. Similarly to the case of flexible

© 2020, esss - all rights reserved


dem technical manual 50
ROCKY

fibers described previously, a joint entity is assumed to be acting at


each one of the contacts between prismatic elements of a flexible Shell.
Forces and moments will be induced in joints as a response to the
relative movement between adjacent elements, as shown schematically
in Figure 3.6.

(a)

Thickness

(b)

Figure 3.6: (a) Portion of a flexible


Shell particle. (b) Forces and moments
Connection face related to a joint acting over an isolated
prismatic element.

Forces and moments at the joints of a Shell particle are calculated


using the same mathematical expressions listed for flexible fibers.
j j
For instance, forces Fn and Fτ are calculated using equations (3.1)
and (3.2), respectively. The relative displacement and velocity vectors
needed in those equations are computed considering the positions
and velocities of the center points of the corresponding rectangular
faces. Moreover, the same decomposition defined in equations (3.3)
and (3.4) is considered for decomposing those vectors in the normal
and tangential directions.
The normal stiffness for the joint is calculated by means of equation
(3.5), considering that the characteristic length for the joint is given

by: 3¯
`j = `t (3.14)
2
where `¯ t is the average length of the triangle edges that define the
shape of the Shell particle. The area A used in equation (3.5) is the
area of the rectangular face associated to the joint, whereas the value

© 2020, esss - all rights reserved


dem technical manual 51
ROCKY

of its Young’s modulus is determined using equation (3.7). On the


other hand, the tangential stiffness for the joint is calculated using the
same expression used for fibers, equation (3.6). In a similar way, the
normal and tangential damping coefficients are calculated by means
of equations (3.8) and (3.9), respectively.
Differently from the flexible fiber case, three components of
moment are computed for a flexible Shell, as depicted schematically
in Figure 3.6. These three components are:

j j
j Kτ In rel Cτ In rel
Mn = θ + ωn (3.15)
A n A
j j
j Kn Iα rel Cn Iα rel
Mα = θ + ωα (3.16)
A α A
j j
j Kn Iβ rel Cn Iβ rel
Mβ = θ + ωβ (3.17)
A β A

where:

• θnrel , θαrel and θ βrel are the relative angular displacements between the
connected elements, around axes coinciding with directions n, α
and β, respectively.

• ωnrel , ωαrel and ω rel


β are the relative angular velocities between the
connected elements, around axes coinciding with directions n, α
and β, respectively.

• In , Iα and Iβ are the principal area moments of inertia of the


rectangular face associated to the joint.

In the previous expressions, n is the same normal direction


considered in the decomposition of the force. Moreover, α and β
are two directions perpendicular to n and orthogonal between them.

3.3 Solid particle shapes


Solid is a category of particle shapes that are fully three-dimensional,
and can be convex or concave. While the category itself is new in
Rocky v4.1, Solid shapes themselves have come standard with Rocky
since its earliest releases. Seven solid shapes are provided by default,
but the possibilities are limitless because custom polyhedral shapes
can be easily imported from STL files. In this version of Rocky,
breakage models can be used with non-sphero, non-meshed Solid
particle shapes.

© 2020, esss - all rights reserved


dem technical manual 52
ROCKY

By default, Solid shapes are considered inflexible objects, however,


polyhedral Solid particles can be simulated also as flexible if they are
composed by multiple bonded elements.5 5
In previous releases of Rocky, particles
composed by Multiple Elements were
referred to as Meshed particles.

3.3.1 Flexible solid particles

A Polyhedron or a Custom Polyhedron is subdivided into tetrahedral


elements when it has to be treated as a flexible particle. Each
tetrahedron composing such particles is connected to an adjacent
element through a triangular face, as depicted in Figure 3.7. Similarly
to the cases of flexible fibers and flexible shells, in order to model
forces and moments between tetrahedral elements, a flexible joint is
considered at each connection. As in those cases, forces and moments
arise from the relative movement between adjacent tetrahedral
elements. Therefore, the modeling of flexible Solid particles in Rocky
is essentially an extension of the model of flexible fibers and shells,
described in previous sections, to fully three-dimensional particles.
The main difference, besides having to deal with higher-dimensional
geometric entities, is the way that stiffnesses are determined for
calculating forces and moments, as described below.

Tetrahedral element (a)

Connection face

(b)

Figure 3.7: (a) Portion of a polyhe-


dral flexible particle decomposed into
tetrahedral elements. (b) Forces and
moments related to a joint acting over a
tetrahedral element.

The normal and shear components of the force at a joint are


calculated using equations (3.1) and (3.2), respectively. The relative
displacement and velocity vectors used in those equations are
computed considering the positions and velocities of the centroids of

© 2020, esss - all rights reserved


dem technical manual 53
ROCKY

the corresponding triangular faces. Again, the same decomposition


defined in equations (3.3) and (3.4) is considered for decomposing
those vectors in the normal and tangential directions.
The elastic properties for the joints between tetrahedral elements
are determined in Rocky from curves obtained by Potapov &
6
Potapov, A. V. and Campbell, C. S.
(1996). A three-dimensional simulation
Campbell.6 Those curves, reproduced in Figure 3.8, relate the of brittle solid fracture. International
Journal of Modern Physics C, 7(5):717–729
dimensionless Young’s modulus Ψ j and the Poisson’s ratio ν j with the
j j j j
stiffness ratio rK = Kτ /Kn for the joints.7 The dimensionless Young’s 7
Here, rK is the stiffness ratio for the
joints, which normally will be different
modulus is defined as: from the contact stiffness ratio consid-
Ej
Ψj = (3.18) ered in some of the tangential force
Kn00 ` T models described in section 2.1.2.

where ` T is a characteristic length of the particle tessellation and Kn00


is the normal stiffness per area unit. In Rocky, ` T is considered equal
to the average length of the edges of all tetrahedra in the particle
tessellation.
The procedure considered to determine the values of the stiffnesses
j j
Kn and Kτ for the joints is as follows:

1. Considering the value of ν j as the Poisson’s ratio of the particle’s


j
material, the corresponding value of the stiffness ratio rK is
determined from curve (b) in Figure 3.8.
j
2. That value of rK is then used to determine the corresponding value
of the dimensionless Young’s modulus Ψ j from curve (a) in Figure
3.8.
j
3. The value of Kn for the joint is determined using the expression:
j
j rE Ep A
Kn = (3.19)
Ψ j `T

where E p is the Young’s modulus of the particle’s material, A is


j
the area of the triangular contacting face, and r E is a user input
parameter in Rocky, listed in the UI as Young’s Modulus Ratio .8 8
This parameter is introduced in order
to allow users to fine-tune the flexibility
j
4. Finally, the value of Kτ is determined using the expression: of the joints. Smaller values of this
parameter will give rise to more flexible
j j j joints, and vice versa. Allowed values
Kτ = r K Kn (3.20) are in the range from 0.01 to 100.

The three components of moment are calculated with the same


expressions listed for the case of flexible shells, that is, equations
(3.15)–(3.17). The relative angular displacements and velocities for the
joints on solid particles are defined in a similar way as in the shell
case. On the other hand, all geometric parameters appearing on those
equations are referred now to the triangular connection faces.

© 2020, esss - all rights reserved


dem technical manual 54
ROCKY

Figure 3.8: (a) Dimensionless Young’s


0.6 modulus and (b) Poisson’s ratio as
a function of the stiffness ratio for
composite bodies formed by Delaunay
tetrahedra. Extracted from Potapov &
0.4 Campbell (1996).

(a)

0.2

0
0 1 2 3 4 5

0.4

0.2

(b)
0

-0.2

0 1 2 3 4 5

Rocky users must be aware that every element forming a flexible


particle will expend nearly the same memory and will demand similar
calculations as a single non-composite particle. As a consequence,
simulations using composite particles will be much more time-
consuming than simulations with similar numbers of single-element
particles. Therefore, users must choose carefully the number
of elements when defining composite flexible particles, avoiding
unnecessary overly-tessellated particles. On the other hand, particles
with flat sides that are discretized by several triangles are especially
difficult for the numerical algorithms in Rocky, because that situation
may lead to multiple simultaneous contacts between two particles or
between a particle and a boundary. When dealing with particles of this
type, the hysteretic linear spring model must be the preferred normal
contact model, since it is less susceptible to numerical problems
caused by multiple contacts.

© 2020, esss - all rights reserved


dem technical manual 55
ROCKY

3.3.2 Convex and concave solid particles

When a Custom Polyhedron shape is imported from a STL file, Rocky


performs a verification check in order to determine if it is convex
or concave. This is important because Rocky will use different
geometrical algorithms for handling the interaction of these particle
shapes depending on whether they are concave or convex. For
concave shapes, all geometric algorithms are much more complex
and, consequently, more demanding in terms of both memory and
processing time.
Figure 3.9 compares two simple polyhedral shapes, one convex
and one concave. Roughly speaking, a concave polyhedron is one
that has some faces forming dents or hollows. In convex shapes, a
segment line joining any two points laying on different non-coplanar
faces is always integrally contained inside the shape, as depicted in
Figure 3.9(a). On the other hand, a similar segment line in a concave
shape can pass through other faces, therefore, it may have portions
outside the shape, as illustrated in Figure 3.9(b).

Figure 3.9: Examples of (a) a convex


shape and (b) a concave shape.

(a) (b)

Rocky uses a systematic verification procedure to determine if an


imported solid shape is concave or convex. In this procedure, a simple
check is made on all triangular faces making up a polyhedral shape.
If the shape is convex, all vertices not associated to a face must lie
on the same side relative to a plane containing the face,9 as in Figure 9
Vertices located on the same plane
as the face are not considered in this
3.10(a). On the other hand, when the shape is concave, vertices will verification.
lie on both sides of the containing plane, as in the example depicted
in Figure 3.10(b). In Rocky, a shape will be marked as concave even if
this configuration arises only once during the referred procedure.

© 2020, esss - all rights reserved


dem technical manual 56
ROCKY

(a)

(b)

Figure 3.10: Possible configurations in


the convexity check considered in Rocky.
In order to determine on which side a vertex is located relative to a
plane, the signed distance from the vertex to the plane is calculated.10 10
An expression for the signed distance
from a point to a plane will give a
If the distances from all vertices to a plane containing a face have positive value if the point is on the same
the same sign, then all of them lie on the same side. A difficulty side as the plane’s normal vector and
negative if it is on the opposite side.
arises for shapes having coplanar faces, because rounding errors can
take some vertices to either side of a plane, causing false positives
for non-convexity. In order to avoid this, a tolerance δ is considered,
so only vertices at a distance |d| > δ are considered located off the
plane.11 11
δ is set in Rocky as 10−6 times the
maximum dimension of the shape. This
When importing a custom polyhedral shape, Rocky will run value was determined heuristically and
automatically the convexity verification described above. If non has proven to work adequately on the
convexity check implemented in Rocky.
convex faces are found, Rocky will mark the shape as concave and,
therefore, the more expensive geometrical algorithms for concave
particles will be used during the simulation, for the corresponding
particle group. If users know that the concavities on the imported
shape are actually very mild or they are in regions of the shape
unlikely to come into contact with other particles, they can manually
mark the shape as non concave in the Shape sub-tab. This will force
Rocky to use the algorithms for convex shapes, however, users must
be aware that this can lead to unexpected results.

© 2020, esss - all rights reserved


dem technical manual 57
ROCKY

4 Collision statistics
Collision statistics modules in Rocky collect relevant data from
collisions occurred during a finite lapse of time, and then associate
the derived statistical information to different entities, in order to
let the users visualize it in useful ways. Such statistical information,
besides being usually more meaningful than simple instantaneous
data, may provide a deeper insight into the dynamics of industrial
particulate processes.
As of Rocky version 4.3, four collision statistics modules are
available in Rocky. The main difference between them is the
spatial scope of the data collection and the way data is associated
to geometrical entities for the purpose of visualization and post-
processing. This is shown schematically in Figure 4.1.
In the first module, boundary collision statistics, collision data is
collected and displayed per boundary triangle. The equivalent module
on the particle side is inter-particle collision statistics, in which data is
collected and displayed per particle. On the other hand, intra-particle
collision statistics only registers collisions involving particles belonging 1
A particle group in Rocky is a category
to a given particle group,1 and associates the data to vertices on its of particles of the same shape and
sharing the same properties.
surface, according to the actual position of the collision points relative

Boundary collision statistics Inter-particle collision statistics Intra-particle collision statistics Inter-group collision statistics

Scope: per particle group pair or


Scope: per boundary triangle Scope: per particle Scope: per vertex/particle group particle group / boundary

Figure 4.1: Four built-in collision statis-


tics modules in Rocky.

© 2020, esss - all rights reserved


dem technical manual 58
ROCKY

to the particle. Finally, inter-group collision statistics gathers collision


data from all collisions involving a pair of particle groups or a particle
group-boundary pair. The specifics of each one of these modules are
described later in this chapter.
The lapse of time in which all collision statistics modules collect
collision data is the interval between two consecutive output times.2 2
An output time is the simulation time
at which Rocky stores relevant data that
Although statistical properties are generated from collision data later can be used for visualization and
registered during a data collection interval, they still will be associated post-processing. The lapse between two
consecutive output times is specified
to an output time for the purpose of visualization and post- through the Output Frequency parameter
processing.3 As indicated schematically in Figure 4.2, if two output in the Solver|Time panel.
3
In order to help clarify if a given prop-
times i and i + 1 delimit the data collection interval i, the statistical
erty is instantaneous or was generated
properties generated from collision data collected in that interval will by a collision statistics module, Rocky
will indicate this in the new Evaluation
be associated to the output time i + 1. column present in every panel listing
boundary or particle properties.
Output time i +1
Output time i

Time

Data collection interval i Figure 4.2: Data collection time interval.

As mentioned previously, each collision statistics module considers


a different geometric entity as a spatial unit for collecting collision
data. This spatial unit is called herein data collection region. For
instance, on boundary collision statistics, this region is the surface of
a triangle making up a boundary. On inter-particle collision statistics,
the data collection region is the whole surface of a particle.

4.1 Collision statistics types


The statistics provided by Rocky can be classified into two main
groups: event-based and integral-type statistics. The specifics of each
one of these groups is described in the following sections.

4.1.1 Event-based statistics

Some collision-related magnitudes can be characterized by a single


value per collision, such as, for instance, the collision duration and
the impact velocity. Rocky can record these magnitudes in order
to compute statistics that allow users to characterize the probability
distributions associated to those magnitudes.

© 2020, esss - all rights reserved


dem technical manual 59
ROCKY

Let ψ be a random variable which have a single observed value


per collision, ψi . Considering that Nc is the number of collisions
occurred during a time collection interval on a data collection region,
the statistics calculated by Rocky based on the collected data are the
following:

N
∑ i=c 1 ψi
Mean value: ψ= (4.1)
Nc
s
N
∑ i=c 1 ( ψi − ψ )2
Standard deviation: σψ = (4.2)
Nc
N
∑ i=c 1 ( ψi − ψ )3
Skewness: Sψ = (4.3)
Nc σψ3
N
∑ i=c 1 ( ψi − ψ )4
Kurtosis: Zψ = (4.4)
Nc σψ4

The values of these statistics can assist users in the characterization


of the probability distribution associated to the random variable ψ.
For instance, the mean is a measure of the central tendency of the
variable for the set of observed values. The standard deviation is a
measure of the dispersion of the observed values in relation to the
mean. Moreover, the value of the skewness indicates how much and
in which direction a distribution departs from symmetry.4 Finally, the 4
Brown, S. (2008–2017). Measures
of shape: skewness and kurtosis.
value of the kurtosis indicates how tall and sharp the central peak of https://brownmath.com/stat/shape.htm
a distribution is.
It is important to specify what is defined as a collision event in the
collision statistics modules available in Rocky. In general, a collision
event occurs whenever a particle’s surface touches the surface of
another particle or a boundary. The collision begins at the instant
that an overlap between two surfaces arises, and finishes when that
overlap decreases to zero.5 5
If immediately after the overlap has
reduced to zero it begins to increase
Particles with complex shapes, such as concave polyhedral or multi- again, this is considered as a new colli-
branched fibers, can touch another particle or boundary at multiple sion event.

locations. As each of these different contacts can begin at a different


time or have a different duration, every one of these contact locations
are considered a separate collision event. Multiple contact locations
can arise also as a consequence of the discretized representation of
the surfaces of particles and boundaries. Such representation includes
triangles, edges, and vertices, and whenever any of those entities
on one surface touches an entity on another surface, an individual
collision event will be registered.

© 2020, esss - all rights reserved


dem technical manual 60
ROCKY

4.1.2 Integral-based statistics

These statistics are based on integral expressions including physical


magnitudes that vary during a collision. Two typical examples are
the statistics based on the work and on the impulse, both of which
are related to the contact forces. Whenever a collision statistics
module requires them, Rocky will compute those integral expressions,
involving the components of the contact forces during a collision.

4.1.2.1 Works

In the case of the work, different definitions may be required for


different statistics. Depending on the actual process being simulated,
some of them may be more useful for a given analysis than others.
Firstly, we can define the work made by the normal contact force
in the loading portion of a collision, that is, the interval during which
the overlap increases. In Rocky this work is called impact work, and is
computed according to:
Z
imp
Wc = Fn dsn (4.5)
c, dsn >0

where Fn is the normal contact force, computed by any of the normal


contact models described in section 2.1.1, while sn is the overlap. In
the normal force-overlap diagram depicted in Figure 4.3, the impact
work is numerically equal to the area below the loading curve OA,
that is, the green area plus the blue area. As the impact work value
will coincide with the maximum energy transferred in a collision, it
is usually considered when analyzing the damage that may lead to
breakage of particles.

L oa d i n g

g
d in
lo a
Un Figure 4.3: Normal force-overlap dia-
O B gram of a simple collision.

Instead of computing the work only during the loading periods,


the computation can extend to the entire collision, including the

© 2020, esss - all rights reserved


dem technical manual 61
ROCKY

unloading periods, until the moment in which the contact ceases.


After the overlap reaches a maximum, the relative movement is
reverted and, consequently, we have dsn < 0 during the unloading
period that follows. Thus, the work computed in the unloading has
the opposite sign with respect to the work done during the loading
period. And consequently, that unloading portion of the work can be
interpreted as recovered mechanical energy, available to the particle
when the collision finishes. The remaining part, which is equivalent
to the green area in Figure 4.3, may be interpreted as the mechanical
energy dissipated in the collision. That work is called dissipation work
in Rocky. Symbolically, the dissipation work related to the normal
contact force is defined as:
Z
Wcdiss, n = Fn dsn (4.6)
c

where the integral is defined on the entire collision c. The dissipation


work computed in that way can be useful when the energy balance
of a process is being analyzed. In cases in which several particles
collide simultaneously, the behavior of the loading-unloading cycles
can be quite complex, with multiple partial unloading and reloading
intervals. In spite of that, the impact work and the dissipation work
are computed always in the way described above. That is, the former
is computed only during the loading intervals, while the latter is
registered during the entire cycle.
Similarly to the dissipation work in the normal direction, Rocky
may calculate also the dissipation work related to the tangential
component of the contact force, defined in an equivalent way:
Z
Wcdiss, τ = Fτ dsτ (4.7)
c

where Fτ is the component of the contact force on the tangential plane,


and sτ is the tangential sliding. In this case the integral is defined
also on the entire collision interval.
Specifically for the case of particle-boundary collisions in which
the boundary is moving, the work made by the contact forces against
the boundary in motion is also calculated. This work is defined by
the expression:
Z Z
Wcb =− b
F · dr = − F · vb dt (4.8)
c c

where F is the total contact force,6 while vb is the boundary velocity at 6


The total contact force is the resultant
of the vector sum of the normal and
the contact point. This work is computed over the entire collision, and tangential contact forces.

© 2020, esss - all rights reserved


dem technical manual 62
ROCKY

can be either positive or negative. It will be positive if the boundary


transfers energy to the particles and negative if the flow of energy is
in the opposite direction.
The values of all integrals defined in this section are approximated
numerically using the trapezoidal rule. Those values are used by the
different collision statistics modules in slightly different ways. The
specific use is indicated later on in the corresponding sections that
describe each collision statistics module individually.

4.1.2.2 Impulses

The computation of forces and/or stresses in the collision statistics


modules require the values of the impulses associated to the contact
forces. The calculation differs depending on the module being
considered. For instance, for the use in the boundary collision
statistics module, the impulse vector acting during an entire collision
is given by: Z
Jc =
c
F dt ≈ ∑ F ∆t (4.9)
c

where F is the total contact force and ∆t is the simulation timestep.


After calculating the impulse vector, its normal and tangential compo-
nents are determined, respectively, using the following expressions:

Jcn = | Jc · n̂ | (4.10)

Jcτ = | Jc − (Jc · n̂) n̂ | (4.11)

where n̂ is, in this case, the unit vector normal to a given boundary
triangle.
For intra-particle and inter-particle collision statistics, the impulses
are calculated directly with the normal and tangential components of
the contact forces. Therefore, we have:
Z
∑ Fγ ∆t
γ
Jc = Fγ dt ≈ for γ = n, τ (4.12)
c c

where Fγ is either the normal or tangential contact force as calculated


using the corresponding normal or tangential model. For the purpose
of computing impulses on particles, every force component is treated
independently as if it were a pure scalar.7 7
The reason for this practice is the
nonexistence of a unique and clearly
defined normal direction for particles,
because the data collection region is
always non-planar in that case.

© 2020, esss - all rights reserved


dem technical manual 63
ROCKY

4.1.2.3 Frequency

Although the definition of the collision frequency is not based on an


integral expression, it is listed here for the sake of completeness. This
frequency is calculated as:
Nc
f = (4.13)
∆tout

where Nc is the total number of collisions registered on the data


collection region, during the data collection time interval.8 Moreover, 8
When determining the value of Nc for
equation (4.13), some effort is made
∆tout is the data collection time interval, which is the time lapse in order to avoid counting collisions
between two consecutive outputs. with multiple contact locations as mul-
tiple collisions, in cases when that
phenomenon is an artifact caused by
the discretized representation of the
surfaces.

4.2 Collision statistics modules


4.2.1 Boundary collision statistics

When the boundary collision statistics module is enabled, data from


all particle-boundary collisions is collected in order to calculate
statistics that will be available for visualization on the surface of
boundaries.9 The data collection units are the boundary triangles 9
Since Rocky allows users to choose
which sub-categories of statistics will be
in this case. This means that a given statistical value displayed effectively calculated, only the collision
on a boundary triangle is representative of data from all collisions data needed for those calculations will
be collected during the simulation, in
involving that triangle that occurred during a data collection interval. order to save memory and disk space.
The available properties in boundary collision statistics are listed
in Table 4.1. When duration, impact velocity or sliding distance
are enabled for calculation, all four statistics will be available for
visualization: mean, standard deviation, skewness, and kurtosis.

Property Type Calculation

Duration event-based eqs. (4.1)–(4.4)


Velocity : impact : normal event-based eqs. (4.1)–(4.4)
Velocity : impact : tangential event-based eqs. (4.1)–(4.4)
Sliding distance event-based eqs. (4.1)–(4.4)
Frequency eq. (4.13)
imp
Intensity : impact integral-based eq. (4.14) using Wcα = Wc from eq. (4.5)
Intensity : dissipation : normal integral-based eq. (4.14) using Wcα = Wcdiss,n from eq. (4.6)
Intensity : dissipation : tangential integral-based eq. (4.14) using Wcα = Wcdiss,τ from eq. (4.7)
Intensity integral-based eq. (4.14) using Wcα = Wcb from eq. (4.8)
Jc = Jcn from eq. (4.10)
γ
Stress : normal integral-based eq. (4.16) using
γ
Stress : tangential integral-based eq. (4.16) using Jc = Jcτ from eq. (4.11)
FEM forces integral-based eqs. (4.17)–(4.18)

Table 4.1: Available boundary collision


statistics, as of Rocky 4.3

© 2020, esss - all rights reserved


dem technical manual 64
ROCKY

Herein, the collision duration is defined as the time period in which an


actual overlap exists between the two colliding entities. On the other
hand, the impact velocity is defined as the relative velocity between
the colliding entities at the instant that the contact begins. Moreover,
the sliding distance is the distance that a particle moves relative to the
boundary during a collision, in a direction parallel to the boundary
triangle where the collision occurs.
In Rocky, intensity is defined as power transferred per unit area. The
basis for the calculation of the intensities are the works made by the
contact forces during a collision, whose calculation was described in
section 4.1.2.1. All forms of intensity provided by boundary collision
statistics are computed per boundary triangle using the following
expression:
N
α ∑ c=c 1 (Wcα )b
Ib,T = (4.14)
A T ∆tout

where A T is the area of the boundary triangle and ∆tout is the data
collection time interval. The sum involves all collisions occurred
during that time interval against the boundary triangle.10 As specified 10
As can be seen from the definition in
equation (4.14), the intensities calculated
in Table 4.1, an intensity will correspond to each work defined in are in fact average values, representative
section 4.1.2.1. An important detail that must be taken into account is of all collisions that occurred during the
data collection time interval within the
that the work used in equation (4.14) is only the part of the collision boundary triangle surface.
work made on the boundary side. This work is considered to be one
half of the whole work made during the collision, that is:11 11
This consideration is made for all
works, except for Wcb , calculated in equa-
1 tion (4.8), which is attributed entirely to
(Wcα )b = 2 Wcα (4.15) the boundary.

The average stress components are computed in a similar way to


intensities. The expression considered for an average stress acting
over a boundary triangle is:
N γ
γ ∑ c=c 1 Jc
σb,T = (4.16)
A T ∆tout
γ
where Jc are the components of the impulse, computed using either
equation (4.10) or equation (4.11), for the normal and tangential
directions, respectively.
Boundary collision statistics is able to record also average collision
forces acting over boundaries. These forces can be exported to a
third-party software to perform structural analysis using the finite
element method (FEM). Usually, boundary conditions for that kind of
analysis are nodal forces, that is, forces acting on the vertices of the
triangles making up the boundaries. Therefore, in order to facilitate
the transfer of data to a finite element software, Rocky provides those

© 2020, esss - all rights reserved


dem technical manual 65
ROCKY

nodal values of forces, based on collected forces per boundary triangle.


These latter ones are determined using the expression:

N γ
γ ∑ c=c 1 Jc
Fb,T = (4.17)
∆tout
γ
where Jc is any of the Cartesian components (γ = x, y, or z) of the
vector impulse computed with equation (4.9).
In order to obtain the nodal values of force, it is considered that
collisions over a boundary triangle are randomly distributed, so the
resulting force on the boundary triangle can be split evenly between
its three vertices.12 Therefore, a component of the combined force on 12
The value attributed to a vertex would
be representative of a region of area
one of those vertices will be given by: equal to one third of the triangle’s area.
NT
1

γ γ
Fb,v = 3 Fb,T (4.18)
T =1
γ
where NT is the number of triangles surrounding a vertex v, while Fb,T
is a component of the average force on every one of those triangles,
obtained using equation (4.17).
FEM forces are the only properties that the boundary collision
statistics module associates only to boundary vertices, for both
visualization and exportation purposes. All other properties are
associated primarily to boundary triangles, although there exists also
the option of displaying them as a node-based field representation
over the boundary surfaces. For this, vertex values are computed as
averages of the values corresponding to the surrounding triangles.

4.2.1.1 Curves

Additionally to the visualization of the spatial distribution of


intensities and stresses over the boundary surfaces, Rocky provides
also time curves in which each point is the integrated value of a given
property over a whole boundary at a given output time. For example,
the integration of an intensity over a boundary will give rise to a
power curve: Z NT
Pbα =
b
I α dA ≈ ∑ α
Ib,T AT (4.19)
T =1

α is a value of intensity computed with equation (4.14) on a


where Ib,T
triangle of area A T located on a boundary b, while NT is the number
of triangles making up that boundary.13 13
If we consider the dissipation intensity
in equation (4.19), we will obtain the
The integration of the stresses over a boundary allows Rocky to dissipation power. Conversely, if we
determine the total forces acting on it. For instance, the normal force consider the impact intensity, we will
obtain the impact power, and so on.

© 2020, esss - all rights reserved


dem technical manual 66
ROCKY

curve is obtained via the integration:

Z NT
Fbn =
b
σn dA ≈ ∑ n
σb,T AT (4.20)
T =1

n is the normal stress computed with equation (4.16) in the


where σb,T
boundary triangle of area A T .
When the Cartesian components of the force are available at the
boundary vertices, because they were calculated through equation
(4.18), the components of the total force over a whole boundary are
obtained using:
Nv
∑ Fb,v
γ γ
Fb = (4.21)
v =1

4.2.2 Intra-particle collision statistics

The intra-particle collision statistics module keeps track of all


collisions involving particles that belong to a given particle group.
Then, depending on the collision location over a particle’s surface, it
associates the collected data to a geometric entity of the discretized
representation of the surface, for the purposes of visualization. Since
all particles in a particle group have the same shape, the statistics
can be displayed over the surface of a single particle of the group.
However, it is important to note that every value displayed at a given
location is representative of all collisions that occurred at that location,
on any particle of the group.
Currently, intra-particle collision statistics is available for poly-
hedral single-sized rigid particles14 and for both flexible fibers and 14
Therefore, particle groups with size
distributions or with a breakage model
flexible solid particles. As illustrated in Figure 4.4, the data collection enabled are excluded from intra-particle
region for polyhedral particles is a non-planar region around a vertex collision statistics.

on the particle surface, while for flexible particles it is the whole

(b)
Data collection region

Figure 4.4: Data collection region for


(a) intra-particle collision statistics on (a)
rigid polyhedral particles and (b) flexi-
ble fibers.

© 2020, esss - all rights reserved


dem technical manual 67
ROCKY

surface of an element making up the particle. In the case of a


polyhedral particle, statistics values are associated to the surface
vertices, therefore, the color displayed at a vertex is related to the
value attributed to that location. Colors at points other than vertices
are determined via interpolation, so the visual representation of
the available statistics usually shows a smooth variation over the
particle’s surface. On the other hand, the statistics calculated for
flexible particles are represented as a single color per element.
It is worth discussing now the criteria considered to associate the
collision data to a vertex in polyhedral particles. The simplest case is
when a collision point coincides with a vertex on the particle’s surface.
Obviously, in that event the data is associated directly to that vertex.
On the other hand, if the contact point is located on an edge, the
association is made to the nearest vertex on that edge. Moreover,
when a collision point is located inside a triangle, a barycentric
subdivision of the triangle is considered in order to determine to
which vertex the collision data must be associated. In this kind of
subdivision, a triangle is divided into three sub-regions of equal
area, each one limited by two line segments whose endpoints are the
triangle centroid and an edge midpoint, as shown in Figure 4.5. If a
collision point lies on one of these sub-regions, then the collision data
is associated to the corresponding vertex.

Vertex

Midpoint

Centroid

Figure 4.5: Construction of a data collec-


tion region on a polyhedral particle.

Table 4.2 lists the available intra-particle collision statistics and the
way they are calculated. The duration and impact velocity statistics
are defined and computed exactly in the same way described in
section 4.2.1, for the boundary collision statistics module. Regarding
the integral-based statistics, there is an important difference that
must be taken into account, however. A value attributed to a vertex
or to an element is the result of the accumulation of collision data

© 2020, esss - all rights reserved


dem technical manual 68
ROCKY

Property Type Calculation

Duration event-based eqs. (4.1)–(4.4)


Velocity : impact : normal event-based eqs. (4.1)–(4.4)
Velocity : impact : tangential event-based eqs. (4.1)–(4.4)
Frequency eq. (4.22)
imp
Intensity : impact integral-based eq. (4.23) using Wcα = Wc from eq. (4.5)
Intensity : dissipation : normal integral-based eq. (4.23) using Wcα = Wcdiss,n from eq. (4.6)
Intensity : dissipation : tangential integral-based eq. (4.23) using Wcα = Wcdiss,τ from eq. (4.7)
γ
Stress : normal integral-based eq. (4.26) using Jc from eq. (4.12), for γ = n
γ
Stress : tangential integral-based eq. (4.26) using Jc from eq. (4.12), for γ = τ

Table 4.2: Available intra-particle colli-


sion statistics, as of Rocky 4.3

coming from, potentially, all particles belonging to the particle group.


Therefore, in order to transform that cumulative value into an average
value representative of a single particle in the group, we need to
divide it by the number of particles that belongs to the group in the
simulation, Npact , at the moment of storing the information. So, for
example, the average collision frequency must be defined as:

Nc
f = (4.22)
∆tout Npact

The average impact and dissipation intensities are calculated using


the expression:
N
α ∑ c=c 1 (Wcα ) p
I p,v = (4.23)
Av ∆tout Npact

where Av is the area of the data collection region around a vertex,15 15


The area of the non-planar region
illustrated in Figure 4.4(a)
for polyhedral particles, or the area of the entire surface of an element,
for flexible particles.
Equation (4.23) takes into account only the work made by the
considered particle. In particle-boundary collisions, the work made
by the normal contact forces is split evenly between the particle and
the boundary, therefore:
1
(Wcα ) p = 2 Wcα (4.24)

On the other hand, in particle-particle collisions, the splitting factor


is considered to be inversely proportional to the stiffness of each
particle:16 16
This comes from the fact that the work
Kn of the normal force on each particle is
(Wcα ) pi = Wα (4.25)
Knpi c proportional to its own normal defor-
mation. In turn, that deformation is
inversely proportional to its stiffness.
where Kn is the equivalent stiffness of the contact, defined similarly
to equation (2.3), and Knpi is the stiffness of the considered particle,

© 2020, esss - all rights reserved


dem technical manual 69
ROCKY

computed as in equation (2.5). Regarding the work related to the


1
tangential force, it is distributed evenly with a 2 splitting factor, as in
equation (4.24).
Finally, the average components of the stress are calculated using
the expression:
N γ
γ ∑ c=c 1 Jc
σp,v = (4.26)
Av ∆tout Npact
γ
where Jc is either the impulse of the normal or the tangential contact
force, calculated as in equation (4.9). The other parameters are the
same described above.

4.2.3 Inter-particle collision statistics

The statistics computed by the inter-particle collision statistics module


involves data collected from collisions that occurred on the entire
surface of a particle. That is, during a data collection interval, all
collisions involving any particle within the simulation will be tracked
and the statistics calculated with the collected data will be associated
to the particle itself. The generated statistics will appear as additional
particle properties in the Rocky UI.17 They can be visualized on a 3D 17
In the Rocky UI, particle properties
generated by the inter-particle collision
view window as any other particle property in Rocky, mapping the statistics module are marked as Statisti-
property value to a color scale. cal, in order to differentiate them from
the built-in particle properties, which
Table 4.3 lists the current available statistics provided by the inter- are marked as Instantaneous in the new
Evaluation column.
particle collision statistics module. All event-based statistics and the
collision frequency are calculated as described for boundary collision
statistics in section 4.2.1, the only difference is that the data comes
from the collisions registered against a particle’s surface during a data
collection interval.

Property Type Calculation

Duration event-based eqs. (4.1)–(4.4)


Velocity : impact : normal event-based eqs. (4.1)–(4.4)
Velocity : impact : tangential event-based eqs. (4.1)–(4.4)
Frequency eq. (4.13)
imp
Power : impact integral-based eq. (4.27) using Wcα = Wc from eq. (4.5)
Power : dissipation : normal integral-based eq. (4.27) using Wcα = Wcdiss,n from eq. (4.6)
Power : dissipation : tangential integral-based eq. (4.27) using Wcα = Wcdiss,τ from eq. (4.7)
γ
Force : normal integral-based eq. (4.28) using Jc from eq. (4.12), for γ = n
γ
Force : tangential integral-based eq. (4.28) using Jc from eq. (4.12), for γ = τ

Table 4.3: Available inter-particle colli-


sion statistics, as of Rocky 4.3

© 2020, esss - all rights reserved


dem technical manual 70
ROCKY

The impact power along with the normal and tangential power are
calculated using the expression:

N
∑ c=c 1 (Wcα ) p
Ppα = (4.27)
∆tout

using the appropriate work, as indicated in Table 4.3. Again, the


sum will involve all registered collisions against the particle’s surface,
during the ∆tout interval. The same considerations made in section
4.2.2 regarding the splitting of the collision works between particles
or particle and boundary applies here.
Similarly, the average normal and tangential forces are computed
by means of: N γ
∑ c Jc
Fp
γ
= c =1 (4.28)
∆tout
γ
in which Jc is the corresponding normal or tangential impulse,
computed as in equation (4.12).

4.2.4 Inter-group collision statistics

The inter-group collision statistics module classifies all collisions that


occurred during a data collection interval, taking into account which
group a colliding particle belongs to,18 and considering all possible 18
Let’s remember again that a particle
group is a category of particles of the
combinations of groups and boundaries. same shape and sharing the same prop-
Let’s illustrate this by means of an example: a simulation that erties.

includes 3 particle groups (P1 , P2 , P3 ), and 2 boundaries (B1 , B2 ). Any


collision in that simulation will involve a pair of particles or a pair
particle-boundary belonging to one of the 12 combinations listed in
Table 4.4.

P1 -P1 P1 -P2 P1 -P3 P1 -B1 P1 -B2 Table 4.4: All possible combinations of 3
particle groups and 2 boundaries.
P2 -P2 P2 -P3 P2 -B1 P2 -B2
P3 -P3 P3 -B1 P3 -B2

For any of such combinations, the inter-group collision statistics


module will generate time curves from data collected in all collisions
involving particles (or a boundary) of the given pair. The curves
generated by this module are listed together with the built-in particle
curves in the Rocky UI.
As of Rocky 4.3, only dissipation energy curves are provided by
the inter-group collision statistics module. Each point in those curves
is obtained by simply summing the dissipation works computed in

© 2020, esss - all rights reserved


dem technical manual 71
ROCKY

all collisions registered involving particles (or a boundary) of the


corresponding pair, during the data collection interval. That is:

Nc
∑ (Wc
diss,γ diss,γ
Epair = )pair (4.29)
c =1

diss,γ
where Wc is the dissipation work per collision, computed either
with equation (4.6), for the normal direction, or equation (4.7), for the
tangential direction.

© 2020, esss - all rights reserved


dem technical manual 72
ROCKY

5 Miscellaneous topics
5.1 Contact detection
Broadly speaking, contact detection comprehends all geometrical
calculations related to the interaction among particles and between
particles and boundaries. Geometrical data needed for the calculation
of contact forces is computed during the contact detection phase.
Since Rocky deals with several types of particle shapes, including
custom polyhedral shapes of arbitrary complexity, contact detection
can easily take up a significant percentage of the total processing
time.
In order to optimize processing time, contact detection operations
are performed in two independent stages, as indicated in the flowchart
shown in Figure 5.1. The first stage is a rough search that aims to
determine a list of the particles that are closest to every particle in the
simulation. Since this is a costly operation, it is performed only after
some number of simulation timesteps, as indicated in the referred
flowchart. In the second stage, which is performed in every timestep,
the exact distances between neighbor particles are computed. In fact,
all relevant geometrical information regarding every pair of neighbor
particle-particle or neighbor particle-boundary is calculated in this
stage, taking into account the most recent positions of the particles in
the simulation.
In Rocky, neighbors of a particle are all particles located at a
distance less than a predetermined value ε, herein called neighboring
distance. This is illustrated schematically in Figure 5.2. The actual
geometry of the particles is not taken into account in the first stage of
neighbor detection. Since this operation must involve every pair of
particles in the simulation, considering their actual geometry would
be infeasible. Instead, in order to determine if a particle is in the
neighborhood region of another, bounding spheres are considered
around every particle.1 Therefore, a particle will be included into the 1
The exception are flexible fibers, since
for them the actual geometry of the
list of neighbors of another if the distance between their bounding constituent sphero-cylinder elements is
spheres is less than ε. considered.

© 2020, esss - all rights reserved


dem technical manual 73
ROCKY

Start Figure 5.1: Contact detection stages in


the general algorithm in Rocky.

First stage: Neighbors detection

New time step

Second stage: Neighbors distance calculation

Contact forces calculation

Motion equations integration

Particles positions update

Yes
Final time step?

No

Need No
neighborhood
Finalupdate?
time level?

End Yes

Figure 5.2: Neighborhood region of a


particle in Rocky.

Base particle

Neighbor particles

Non-neighbor particles

Neighborhood region

© 2020, esss - all rights reserved


dem technical manual 74
ROCKY

In the second stage of contact detection, neighbor pairs of particle-


particle or particle-boundary are examined in detail for determining
the relevant geometric parameters needed in the physical contact
models. Here, the actual geometries of the particles are considered,
therefore, the complexity of the calculations will depend on the type
of particles employed in the simulation. By far, the simplest case is
the contact between spherical particles. On the other hand, contact
calculations between polyhedral particles, as those depicted in Figure
5.3, are more difficult, since multiple scenarios must be examined. For
instance, the contact between polyhedra might be vertex-to-vertex,
vertex-to-edge, vertex-to-face, edge-to-edge, edge-to-face, or even face-
to-face. An extra level of complexity is added when the particles are
concave, since in this case multiple contacts can arise between the
same pair of particles.

Figure 5.3: Exact distance calculation


n
between neighbor particles.

Contact plane

(b)

(a)

Regardless of the actual complexity of the algorithms, the end


result is the set of all relevant geometrical parameters associated to a
contact, such as the distance between particles, the application point
of the contact forces, the orientation of the normal to the contact, etc.
Physical contact will arise whenever the calculated distance between
two particles or between a particle and a boundary is negative. The
distance in this case is the normal overlap, needed for the calculation
of the normal contact force.
After some number of timesteps, the list of neighbors for all parti-
cles in the simulation must be updated. Otherwise, some collisions
between non-neighbor particles could arise and the simulation would
not be able to detect them. Therefore, as illustrated in Figure 5.4, in
order to prevent missing collisions, the maximum distance that any
pair of non-neighbor particles can move relative to each other is ε, the

© 2020, esss - all rights reserved


dem technical manual 75
ROCKY

neighboring distance. Whenever that distance is reached, the lists of


neighbors of all particles in the simulation must be updated. Since
the computation of the exact distance can be expensive, a simplified
approximate upper bound is computed in Rocky, which guarantees
that no collision will be missed in the simulations.

Figure 5.4: Before updating neighbor


Non-neighbor particles lists, the maximum distance that any
pair of non-neighbor particles can move
relative to each other is ε.

The neighboring distance ε plays an important role on the


performance of the contact detection algorithms in Rocky. Figure
5.5 illustrates two opposite cases regarding the magnitude of ε. In
the first case, when the neighboring distance is low, the size of the
neighbors lists will be small and, therefore, the processing time in
the second stage of detection will be also low. This is because only
a few pairs of particles will need to be processed. However, this
situation will demand more frequent updates of the neighbor lists,

(a) (b)

Figure 5.5: Two cases with different


magnitude of the neighboring distance
ε.

© 2020, esss - all rights reserved


dem technical manual 76
ROCKY

and, consequently the total processing time of the first stage of contact
detection may be high. On the other hand, in the situation depicted
in Figure 5.5(b), when the value of ε is relatively large, the opposite
behavior is usually observed. In this case, the elevation in the number
of neighbors per particle will produce a significant increase in the
processing time of the second stage. But, at the same, less frequent
updates of the neighbor lists will be needed, therefore, the amount of
processing time of the first stage will decrease.
Users can override the default value of ε determined internally in
Rocky. This parameter can be specified in the Advanced sub-tab of the
Solver panel,2 where it is listed as Neighboring Distance . As ε can have a 2
See also the "About the Solver Parame-
ters" topic in the Rocky User Manual.
huge impact on the total processing time of a simulation, users must
be very careful when specifying its value.

5.2 Timestep calculations


The timestep is one of the main factors that determine the amount
of time required for the completion of a simulation in Rocky. The
simulation timestep should be large enough to ensure reasonable
completion time, but small enough to guarantee the accuracy and
stability of the simulation.
It is a customary practice in DEM to set the timestep size as
a fraction of the oscillation period of a equivalent mass-spring

system,3 , 4 2π m/K, where m and K are critical values of mass 3
Malone, K. F. and Xu, B. H. (2008).
Determination of contact parameters for
and stiffness, respectively. That period of time will approximate discrete element method simulations of
the duration of the shortest possible collision among all possible granular systems. Particuology, 6:521–
528
combinations of particles and boundaries. The timestep value is 4
Tsuji, Y., Kawaguchi, T., and Tanaka, T.
chosen in a way that even that shortest collision may be solved (1993). Discrete particle simulation of
two-dimensional fluidized bed. Powder
numerically with a reasonable resolution. Technology, 77:79–87
In Rocky, the timestep value is calculated automatically at the
beginning of a simulation, using the expressions listed below for
each normal contact model. If necessary, however, users can override
that value and set a new one determined by their own means. In
order to do so, the Fixed Timestep option must be turned on from the
Advanced sub-tab of the Solver panel.5 Enabling this option allows 5
The Advanced sub-tab is visible only
when the Advanced Features option is
the subsequent Timestep value to be defined. Users must be aware, enabled. For more details about this,
however, that the timestep value is a critical parameter in Rocky, so please refer to the "About Setting Global
Preferences" section in the Rocky User
its manual modification is not recommended in general, because it Manual.
may lead to either instability of the solution process or excessively
long simulation times.

© 2020, esss - all rights reserved


dem technical manual 77
ROCKY

5.2.1 Timestep for the hysteretic linear spring model

In the hysteretic linear spring model, the loading and the unloading
stages of the contact occur with two different values of stiffness, Knl
and Knu , respectively.6 Therefore, the calculation of the timestep takes 6
See section 2.1.1.1 for more details
about the definition of these parameters.
into account those two stages independently. That calculation can be
summarized into the following expression:
s !
m∗ π m∗
r
π
∆t = min l
, (5.1)
2N∆t Knl 8 Knu
where:

• m∗ is the effective mass, defined in equation (2.10).


l is the minimum number of timesteps per loading cycle, which
• N∆t
is a user input parameter listed as Loading N-steps in the Rocky UI.7 7
The default value for N∆t l in Rocky

is 15. To ensure the accuracy of the


calculations, it is recommended that the
For using equation (5.1), all particle-particle and particle-boundary
value of this parameter be at least 10.
possible combinations are examined, searching for the critical values
√ √
of m∗ /Knl and m∗ /Knu . That equation guarantees that, during
l
any collision, the loading portion will be discretized with at least N∆t
timesteps, while the unloading will comprehend at least 4 timesteps.

5.2.2 Timestep for the linear spring-dashpot model

In the linear spring-dashpot model, the contact stiffness for the


loading and the unloading have the same value. Hence, it is enough
to analyze the loading for determining the timestep when that model
is used. The expression considered for that is simply:
s !
π m∗
∆t = l
min (5.2)
2N∆t Knl

where all parameters have the same definition as in equation (5.1). All
particle-particle and particle-boundary possible contacts are examined

for finding the minimum value of m∗ /Knl .

5.2.3 Timestep for the Hertzian spring-dashpot model

The Hertzian spring-dashpot model is a nonlinear model, so the


contact stiffness depends on the normal overlap sn . The expression
for the contact stiffness on this model is:8 8
For the definition of the parameters on
this expression, please refer to section
√ 1 2.1.1.3.
K H = 34 E∗ R∗ sn2 (5.3)

© 2020, esss - all rights reserved


dem technical manual 78
ROCKY

When the Hertzian spring-dashpot model is used, the timestep is


determined by means of:

m∗
r 
π
∆t = l
min (5.4)
2N∆t KH

in which K H is estimated considering that the magnitude of the


maximum overlap is 10% of the effective radius R∗ . Again, all particle-
particle and particle-boundary possible combinations are examined

for finding the minimum value of m∗ /K H .

5.2.4 Numerical softening factor

As previously mentioned, the size of the timestep is one of the main


factors that determine how much processing time a simulation will
demand. It would be desirable to increase the timestep as much
as possible, without compromising stability or accuracy, in order to
speed up a simulation.
As described in previous sections, the timestep is always inversely
proportional to the square root of a critical stiffness. Because of this,
one of the techniques most frequently used in order to speed up a
DEM simulation is the reduction of the stiffness.9 Rocky allows the 9
Lommen, S., Schott, D., and Lodewijks,
G. (2014). DEM speedup: stiffness
users to use this technique through a global reduction factor that effects on behaviour of bulk material.
multiplies all stiffness values computed in the different models. This Particuology, 12:107–112

parameter is listed in the Rocky UI as Numerical Softening Factor , which


is located on the Momentum tab of the Physics panel.
As an example, reducing the value of the numerical softening
factor from 1 to 0.01 will increase the timestep approximately 10
times. This change may speed up the simulation in that proportion
also. However, users must be very careful when specifying values
for the numerical softening factor. Values that are too low can lead
to very large overlaps, and therefore, serious stability and accuracy
issues in a simulation.

5.3 Sieve size calculation


The sieve size of a particle is defined as the dimension of the smallest
square aperture through which the particle can pass, as depicted in
Figure 5.6. For polyhedral and custom solid particle shapes, Rocky
uses an approximate method to estimate the sieve size.

© 2020, esss - all rights reserved


dem technical manual 79
ROCKY

Major axis Figure 5.6: Sieve size of a polyhedral


particle.

As a first step, the major axis of the particle is determined. In order


to do this, all possible pairs of vertices on the particle’s geometry are
considered, looking for the two most distant vertices. In the algorithm
implemented in Rocky, it is considered that the major axis of the
particle will pass through those two vertices.
In the second stage of the algorithm, the largest particle dimension
perpendicular to the major axis is sought. This will be the sieve size
of the particle. In order to determine that, a sequence of slices of the
particle’s geometry is considered. As depicted in Figure 5.7, each one
of those slices is limited by two planes orthogonal to the major axis
passing through the end points of an edge.

Figure 5.7: Example of a slice considered


for determining sieve size candidates.

Edge i

Then, all vertices contained between the two limiting planes are
projected orthogonally onto any of those planes. The next step will

© 2020, esss - all rights reserved


dem technical manual 80
ROCKY

be to determine the sieve size candidate associated to the slice. In


order to do that, the distances between the projections of both vertices
at the ends of the edge i and the projections of all other vertices are
calculated. The sieve size candidate associated to the slice i will be
the largest of those distances, dmax,i , as depicted in Figure 5.8.

Figure 5.8: Projection of the vertices onto


an orthogonal plane to the major axis.

Edge i

In the end, the estimated sieve size will be the maximum value
among the sieve size candidates determined considering the slices
associated to all edges on the particle’s geometry. That is:

Ls = max dmax,i (5.5)
i

The described algorithm works well for irregular particles, for


instance, those representing rocks or rock fragments. For regular
simple shapes, such as cubes or parallelepipeds, the algorithm usually
will overestimate the sieve size.

© 2020, esss - all rights reserved


dem technical manual 81
ROCKY

6 Bibliography
Ai, J., Chen, J. F., Rotter, J. M., and Ooi, J. Y. (2010). Assessment of
rolling resistance models in discrete element simulations. Powder
Technology, 206:269–282.

Antypov, D. and Elliott, J. A. (2011). On an analytical solution for the


damped hertzian spring. Europhysics Letters, 94(5).

Archard, J. F. (1980). Wear theory and mechanisms. In Wear control


handbook. American Society of Mechanical Engineers.

Barrios, G. K. P., Carvalho, R. M., and Tavares, L. M. (2011). Modeling


breakage of monodispersed particles in unconfined beds. Minerals
Engineering, 24:308–318.

Batchelor, G. K. and O’Brien, R. W. (1977). Thermal or electrical


conduction through a granular material. Proc. R. Soc. Lond. A.,
355:313–333.

Bierwisch, C., Kraft, T., Riedel, H., and Moseler, M. (2009). Three-
dimensional discrete element models for the granular statics and
dynamics of powders in cavity filling. Journal of the Mechanics and
Physics of Solids, 6257:10–31.

Brown, S. (2008–2017). Measures of shape: skewness and kurtosis.


https://brownmath.com/stat/shape.htm.

Carvalho, R. M. and Tavares, L. M. (2013). Predicting the effect


of operating and design variables on breakage rates using the
mechanistic ball mill model. Minerals and Engineering, 43:91–101.

Cundall, P. A. and Strack, D. L. (1979). A discrete numerical model


for granular assemblies. Geotechnique, 29(1):47–65.

Guo, Y., C., W., Curtis, J. S., and Xu, D. (2018). A bonded sphero-
cylinder model for the discrete element simulation of elastic-plastic
fibers. Chemical Engineering Science, 175:118–129.

Hertz, H. (1882). Über die berührung fester elastischer körper (On the
contact of elastic solids). J. Reine Angewandte Mathematik, 92:156–171.
English translation, Macmillan, London, 1896.

© 2020, esss - all rights reserved


dem technical manual 82
ROCKY

Johnson, K. L. (1985). Contact mechanics. Cambridge University Press.

Johnson, K. L., Kendal, K., and Roberts, A. D. (1971). Surface energy


and the contact of elastic solids. Proceedings of the Royal Society A:
Mathematical, Physical and Engineering Sciences, 324:301–313.

Lommen, S., Schott, D., and Lodewijks, G. (2014). DEM speedup:


stiffness effects on behaviour of bulk material. Particuology, 12:107–
112.

Malone, K. F. and Xu, B. H. (2008). Determination of contact


parameters for discrete element method simulations of granular
systems. Particuology, 6:521–528.

Morris, A. B., Pannala, S., Ma, Z., and Hrenya, C. M. (2016).


Development of soft-sphere contact models for thermal heat
conduction in granular flows. AIChE Journal, 62(12):4526–4535.

Pasha, M. (2013). Modelling of flowability measurement of cohesive powders


using small quantities. PhD thesis, University of Leeds, UK.

Pasha, M., Dogbe, S., Hare, C., Hassanpour, A., and Ghadiri, M. (2014).
A linear model of elasto-plastic and adhesive contact deformation.
Granular Matter, 16:151–162.

Potapov, A. and Donahue, T. (2012). Computer simulation of coal


breakage in conveyor transfer chutes with Rocky discrete element
method package. Technical report, Rocky DEM, Inc.

Potapov, A. V. and Campbell, C. S. (1996). A three-dimensional


simulation of brittle solid fracture. International Journal of Modern
Physics C, 7(5):717–729.

Qiu, X., Potapov, A., Song, M., and Nordell, L. (2001). Prediction of
wear of mill lifters using discrete element method. In 2001 SAG
Conference Proceedings.

Radl, S., Radeke, C., Khinast, J. G., and Sundaresan, S. (2011).


Parcel-based approach for the simulation of gas-particle flows. In
Proceedings of the 8th international Conference on CFD in the Oil & Gas,
Metallurgical and Process Industries., pages 124–134.

Schwager, T. and Pöschel, T. (2007). Coefficient of restitution and


linear-dashpot model revisited. Granular Matter, 9:465–469.

Shi, F. and Kojovic, T. (2007). Validation of a model for impact


breakage incorporating particle size effect. International Journal of
Mineral Processing, 82(3):156–163.

© 2020, esss - all rights reserved


dem technical manual 83
ROCKY

Tavares, L. and King, R. (1998). Single-particle fracture under impact


loading. International Journal of Mineral Processing, 54:1–28.

Tavares, L. M. (2009). Analysis of particle fracture by repeated


stressing as damage accumulation. Powder Technology, 90(3):327–339.

Tavares, L. M. and Carvalho, R. M. (2009). Modeling breakage rates


of coarse particles in ball mills. Mineral Engineering, 22:650–659.

Tavares, L. M. and King, R. P. (2002). Modeling of particle fracture


by repeated impacts using continuum damage mechanics. Powder
Technology, 123(2):138–146.

Tsuji, Y., Kawaguchi, T., and Tanaka, T. (1993). Discrete particle


simulation of two-dimensional fluidized bed. Powder Technology,
77:79–87.

Tsuji, Y., Tanaka, T., and Ishida, T. (1992). Lagrangian numerical


simulation of plug flow of cohesionless particles in a horizontal
pipe. Powder Technology, 71:239–250.

Vargas, W. L. and McCarthy, J. J. (2001). Heat conduction in granular


materials. AIChE Journal, 47(5):1052–1059.

Vogel, L. and Peukert, W. (2005). From single particle impact


behaviour to modelling of impact mills. Chemical Engineering Science,
60(18):5164–5176.

Walton, O. R. and Braun, R. L. (1986). Viscosity, granular-temperature,


and stress calculations for shearing assemblies of inelastic, frictional
disks. Journal of Rheology, 30:948–980.

Wensrich, C. M. and Katterfeld, A. (2012). Rolling friction as a


technique for modelling particle shape in DEM. Powder Technology,
217:409–417.

© 2020, esss - all rights reserved

You might also like