Вы находитесь на странице: 1из 26

INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN ENGINEERING

Int. J. Numer. Meth. Engng 2007; 70:379404


Published online 17 October 2006 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/nme.1881

Discrete element method for modelling solid


and particulate materials

Federico A. Tavarez and Michael E. Plesha,


Department of Engineering Physics, University of Wisconsin-Madison, Madison, WI, U.S.A.

SUMMARY
The discrete element method (DEM) is developed in this study as a general and robust technique for
unified two-dimensional modelling of the mechanical behaviour of solid and particulate materials, including
the transition from solid phase to particulate phase. Inter-element parameters (contact stiffnesses and
failure criteria) are theoretically established as functions of element size and commonly accepted material
parameters including Youngs modulus, Poissons ratio, ultimate tensile strength, and fracture toughness.
A main feature of such an approach is that it promises to provide convergence with refinement of a
DEM discretization. Regarding contact failure, an energy criterion based on the materials ultimate tensile
strength and fracture toughness is developed to limit the maximum contact forces and inter-element
relative displacement. This paper also addresses the issue of numerical stability in DEM computations
and provides a theoretical method for the determination of a stable time-step. The method developed
herein is validated by modelling several test problems having analytic solutions and results show that
indeed convergence is obtained. Moreover, a very good agreement with the theoretical results is obtained
in both elastic behaviour and fracture. An example application of the method to high-speed penetration
of a concrete beam is also given. Copyright q 2006 John Wiley & Sons, Ltd.

Received 11 April 2006; Revised 25 July 2006; Accepted 2 August 2006

KEY WORDS: discrete element method; particulate media; solid media; clusters; convergence; fracture
energy

Correspondence to: Michael E. Plesha, Department of Engineering Physics, University of Wisconsin-Madison, 1500
Engineering Drive, Madison, WI 53706, U.S.A.

E-mail: plesha@engr.wisc.edu

Present address: ExxonMobil Upstream Research Company, P.O. Box 2189, Houston, TX 77252, U.S.A.

Contract/grant sponsor: National Highway Institute


Contract/grant sponsor: U.S. Air Force Office of Scientific Research (AFOSR); contract/grant number: F49620-
03-0216

Copyright q 2006 John Wiley & Sons, Ltd.


380 F. A. TAVAREZ AND M. E. PLESHA

1. INTRODUCTION

The fracture behaviour of solids such as concrete, rock, ceramics, and other brittle materials
under severe loading has become an area of increased research in recent years. These materials
are complex and extremely heterogeneous, especially after they degrade from solid to particulate.
Reproducing the behaviour exhibited by these materials with continuum methods requires complex
constitutive models containing a large number of parameters and/or internal variables, such as
yield surfaces and equation of state descriptors. The discrete element method (DEM) originally
developed by Cundall and Strack [1] has proven to be a powerful and versatile numerical tool
for modelling the behaviour of granular and particulate systems [211], and also for studying the
micromechanics of materials such as soil at the particle level. However, the method also has the
potential to be an effective tool to model continuum problems (i.e. solids), especially those that
are characterized by a transformation from a continuum to a discontinuum. Such problems include
failure of concrete structures, fragmentation of rock due to blasting, and fracture of ceramics and
other quasi-brittle materials under high velocity impact.
For problems with severe damage, DEM offers a number of attractive features over continuum
based numerical methods, with the primary feature being a seamless transition from solid phase
to particulate phase. Continuum based methods such as the finite element method are challenging
to apply to these problems and are plagued by the need for continuum constitutive models, severe
element distortion, and frequent re-meshing. In contrast, DEM has the capability to capture the
complicated behaviour of actual materials by using a discretization scheme that is simple in concept
and implementation, and with simple assumptions and parameters that govern the micro level
elementelement interactions. Complex macroscopic behaviour such as strain softening, dilation,
and fracture arise automatically from extensive microcracking in the DEM medium.
The present study develops the DEM as a general, robust, and scalable computer technique for
unified modelling of the mechanical behaviour of solid and particulate materials, including the
transition from solid phase to particulate phase. Most of the past research in DEM for modelling
continuum behaviour and fracture rely on calibration processes to determine the correct inter-
element parameters for a specific problem [1215]. Consequently, the results that are produced
are likely to be DEM element size dependent. In this work, we attempt to theoretically determine
such parameters as a function of known material properties and DEM element size. A weakness
of DEM is that its convergence properties are not understood. The crucial question is whether
convergence (in both elastic behaviour and material failure) is obtained as DEM element size
vanishes in the limit of model refinement. Therefore, another goal of our investigation is to study
DEM convergence for modelling elastic behaviour and fracture of solids.

2. THE DISCRETE ELEMENT METHOD (DEM)

The DEM discretizes a material using rigid elements of simple shape that interact with neighbouring
elements according to interaction laws that are applied at points of contact. The rigid elements can
have a variety of shapes, although most often they are circular or spherical due to the simplicity
and speed of the contact detection algorithm. The analysis procedure consists of three major
computational steps: internal force evaluation, in which contact forces are calculated; integration
of equations of motion, in which element displacements are computed; and contact detection, where
new contacts are identified and broken contacts are removed. In a DEM analysis, the interaction of

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 381

the elements is treated as a dynamic process that alternates between the application of Newtons
second law and the evaluation of a forcedisplacement law at the contacts. Newtons second law
gives the acceleration of an element resulting from the forces acting on it, including gravitational
forces, external forces prescribed by boundary conditions, and internal forces developed at inter-
element contacts. The acceleration is then integrated to obtain the velocity and displacement. The
forcedisplacement law is used to find contact forces from known displacements. The equations
of motion are integrated in time using the central difference method. Details of this process are
given in [1]. The method can be computationally very demanding and thus, efficient algorithms,
especially for the internal force evaluations and contact detection, must be used. Computational
effectiveness will be particularly important for three-dimensional discretizations, the use of which
is inevitable for obtaining fully realistic and accurate models for many applications.

3. DEM FOR MODELLING PARTICULATE MATERIALS

It is well known that the mechanical behaviour of sands and other granular cohesionless geoma-
terials is different in important respects from that of other materials, and that this difference is
due to the particulate nature of the medium. As such, the DEM has become accepted and widely
used to model the mechanical behaviour and flow of particulate geomaterials. An extension to
industrial geomaterial applications has occurred over the past several years with modelling of ball
mills, dragline excavators, and material transport using conveyor belts. A largely separate research
activity to model powders and materials processing is also ongoing.
The present work uses a modified version of the program TRUBAL [16, 17]. Since 1978, many
researchers have adapted this program to solve specific problems with emphasis on applications
to granular materials. Soil is inherently discontinuous and can be effectively modelled by the
DEM. In terms of computational convenience, two-dimensional circular disk elements and three-
dimensional spherical elements are ideal choices for DEM element shape. However, in some
applications, such as flow problems, the medium may not be adequately represented by using
such smooth elements due to excessive element rotations that occur, and often the agreement
between the shear strength of the DEM medium and the natural material being modelled is
poor. In order to improve agreement, some researchers have employed the ad hoc practice of
artificially constraining the rotations of elements periodically throughout the course of a simulation.
A better approach is to impart some shape complexity to the particles being modelled in a DEM
discretization, and to this end, several investigators have developed DEM elements having more
complex shape than circular or spherical. For instance, Ghaboussi and Barbosa [2] developed three-
dimensional DEM elements consisting of rigid six-sided polygons that can interact with each other
through different types of contacts, and Ting [11] developed DEM elements having elliptical shape.
A contact algorithm for DEM elements of arbitrary geometries was developed by Williams et al.
[18]. Another technique to improve the accuracy of DEM models while retaining the simplicity and
speed of the contact detection algorithms for simple-shape elements was introduced by Jensen et al.
[6, 7], where particles are modelled by combining several smaller DEM elements of simple shape
into clusters that act as a single, larger rough particle. In their research, DEM clusters were formed
by bonding a number of circular-shaped elements into a semi-rigid configuration. In simulations
of media-structure interface shear tests, their results indicate that clustering gives rise to lower
element rotation and increased shear resistance of the medium that is in better agreement with
natural materials. Another form of clustering has been reported in which displacement constraints

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
382 F. A. TAVAREZ AND M. E. PLESHA

are used to render a cluster to be truly rigid [8]. While this approach can be more efficient, primarily
because internal force computations for the intra-cluster contacts is not performed, knowledge of
these intra-cluster forces is desirable for purposes of implementing cluster failure mechanisms, as
is discussed next.
A very significant feature of the clustering developed in [6, 7] is the opportunity to easily
incorporate particle damage such as grain crushing, wear of roughness, etc. Indeed, it appears
that grain crushing is an important result, or initiator, of shear zone formation in sand [9]. Jensen
et al. [10] used an energy density criterion to determine if enough work has been accumulated by a
DEM element of a cluster to warrant its separation (breakage) from the cluster. During the course
of computation, the increment of sliding work done on a particular DEM element i is computed
at each time step by summing the work increments for all extra-cluster contacts for the element

dWi = Ft dst (1)
number of
extra-cluster
contacts

where Ft is the tangential force at the contact and dst is the increment of relative tangential plastic
(sliding) displacement between the DEM element and a contacting neighbour. The total work is
then obtained by integrating (summing) the work increments

Wi = dWi (2)

Therefore, an element separates from its cluster once the total work Wi exceeds a maximum
Wimax , computed by the product of a user-specified critical energy density W0 and the DEM
element volume Vi .
An important conclusion to be drawn here is that it appears that accurate modelling of a
particulate material, such as sand, requires that individual particles be modelled as solids, with due
respect for the their ability to undergo damage due to wear, and damage due to fracturing. The
work of Jensen et al. [10] presents reasonable criteria for modelling damage due to wear, which is
a more or less gradual process that depends on frictional sliding. The work reported in subsequent
sections of this paper will allow damage due to grain fracture, which should enhance the ability
of DEM to model particulate materials.

4. DEM FOR MODELLING SOLID MATERIALS

In our DEM approach to modelling solids, we expand on the clustering concept introduced by
Jensen et al. [6, 7, 10] to use megaclusters, wherein the entire volume of a solid is modelled by
bonding individual DEM elements, or clusters of DEM elements, together. For a solid with no
initial cracks (i.e. a nominally homogeneous material), all contacts in the DEM discretization
are initially bonded with the behaviour that is schematically illustrated in Figure 1. During the
course of a simulation, if a bonded contact fails, according to criteria to be described, the contact
becomes frictional as also shown in Figure 1, if indeed the contact still exists (i.e. if the two
DEM elements are still being pressed into one another), or the contact is destroyed if the two
elements have separated. Thus, for a bonded contact, the normal direction spring acts in tension and
compression, with a relationship between normal force and normal relative displacement given by

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 383

Kt Ct Kt Ct

Cn Cn

Kn Kn

bonded contact frictional contact

Figure 1. Discrete element interaction models for bonded and frictional contacts.

Figure 2. Generation of a cantilever beam model by consolidating three-element DEM clusters.

the elastic stiffness K n , and in the tangential direction, changes of shear force are linearly related
to changes of shear relative displacement by a stiffness K t . For frictional contact, the normal
direction spring acts only in compression and the shear force is limited by a Coulomb friction
law. Consequently, when the computed shear force reaches this maximum value, the inter-element
contact undergoes sliding. The computer implementation uses Rayleigh or proportional damping
[19], and therefore damping parameters are chosen as fractions of critical damping at two desired
frequencies.
Figure 2 shows an example of the creation of a DEM model of a cantilever beam. In the
generation of this model, a domain is initially populated by a large number of clusters (three-
element clusters are used in Figure 2), which are then compressed or consolidated into the shape
of the beam. After consolidation, the three-element clusters that make contact are bonded yielding
a single megacluster that has the desired shape. After this step, all contacts (intra-cluster and inter-
cluster) are treated as inter-cluster contacts. The main purpose of using clusters in the consolidation
process is to add irregularity and randomness to the discretization, and to avoid having large regions
being consolidated into a close-packed arrangement. For circular or spherical elements, as shown
in Figure 3(a), a frictional contact is detected when
dR A + R B (3)
Due to the consolidation process, adjacent elements might slightly overlap, or may be slightly
separated, as shown in Figure 3(b). In our algorithm to determine which contacts should be bonded,
we prescribe a search radius rs around each element, so that if the surface of any other element is

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
384 F. A. TAVAREZ AND M. E. PLESHA

initial element
separation
d Rb

element rs
overlap
Ra

x2

x1
(a) (b)

Figure 3. Frictional contact detection and initial bonded contacts.

found within this radius, the element will be initially bonded to that neighbour. For this study, the
value of rs was chosen rather arbitrarily (rs Ri + Ri /10, where Ri is the element radius), since
the main purpose of this parameter is to ensure a dense bond network in the medium. In order
to have zero initial contact forces throughout the megacluster after consolidation is completed,
element displacement interactions are measured relative to the initial distance between bonded
elements. As such, the packing scheme and wall forces used for the consolidation process are not
likely to significantly influence the bulk macroscopic properties of the model.
Once the models geometry is established, boundary conditions are prescribed, such as the
built-in support at the left-hand side shown in Figure 2. Note that the medium that is produced has
some irregularity and porosity that is typical of a material such as concrete. As such, this method
does not rely upon any assumptions about a distribution of initial flaws in the medium. Also, this
generation scheme allows for molding of structures or structural components of arbitrary shape.
Further, for modelling higher porosity solids, DEM elements may be randomly removed until the
desired porosity is obtained. It should be noted that although the true microstructural properties
of a material such as concrete could in principle be represented using this method, this was not
attempted in the present study, and therefore the DEM medium generation is not being done in a
controlled fashion to model the actual microstructure of a material such as concrete.

5. DETERMINATION OF INTER-ELEMENT NORMAL AND SHEAR STIFFNESSES

The macroscopic linear elastic behaviour of isotropic materials can be characterized by two elastic
constants: Youngs modulus E and Poissons ratio . The model parameters governing DEM element
interactions (i.e. K n and K t ) are typically determined by the ad hoc process of validating a numerical
simulation of a standard laboratory test with an actual experimental result [12, 13]. However, such
results are highly likely to depend on DEM element size. That is, if a new discretization for a
particular problem is produced but with different DEM element size (such as making the element
size smaller to improve accuracy of the simulation), the calibration process would need to be
repeated in its entirety to obtain the new DEM parameters.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 385

4 3 4 3
isotropic
material
h 5 1 2 2 3R 5 1 2
element

(thickness = t)
6 7 6 7

(a) l (b) 4R (c)

Figure 4. Close-packed DEM unit cell for determination of inter-element spring constants.

In the present study we use a new approach wherein we attempt to theoretically establish the inter-
element normal and tangential stiffnesses (and failure parameters, to be discussed later) as functions
of element size and commonly accepted material parameters including Youngs modulus and
Poissons ratio. This is done by developing a DEM model of a unit cell of material. Figure 4(a) shows
an isotropic solid material element (with known elastic modulus and Poissons ratio) subjected to
uniaxial stress. The volume of material is then modelled using the DEM close-packed unit cell
shown in Figure 4(b). The unit cell contains seven elements having three degrees of freedom (d.o.f.)
per element (two translations and one rotation). Due to the symmetry of loading, all rotations in
the unit cell are zero. Therefore, the matrix equation for the 14 translational d.o.f. can be expressed
as

[K]{d} = {f} (4)

where the contents of the matrix [K] are given in Reference [20]. When subjected to loading shown
in Figure 4(a), the material element will undergo rigid body motion and uniform straining. The rigid
body motion is irrelevant for our purposes, and thus we impose displacement constraints on the
unit cell of material shown in Figure 4(b) that describes a uniform straining deformation, including
the Poisson effect. For a uniform straining deformation, such constraints can be implemented by
expressing the 14 d.o.f. unit cell displacement vector as a function of only 2 d.o.f. (horizontal
displacement of element 1 and vertical displacement of element 3, shown in Figure 4). That is,

{d} = [T]{d } (5)

where [T] is a transformation matrix and {d } is the displacement vector containing the two
retained d.o.f. (displacements are measured relative to element 5). Substituting Equation (5) into
Equation (4) and premultiplying by [T]T provides

[T]T [K][T]{d } = [T]T {f} (6)

Therefore, the reduced system can be expressed as

[K ]{d } = {f } (7)

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
386 F. A. TAVAREZ AND M. E. PLESHA

where [K ] = [T]T [K][T] and {f } = [T]T {f}. Assuming small deformations, the reduced system
in Equation (7) is
   
2K n c2 + 2K t s 2 + 4K n 4K n cs 4K t cs dx1 8Rsx x t
= (8)
4K n cs 4K t cs 8K n s 2 + 8K t c2 d y3 0

where R is the DEM element radius, t is the thickness (out-of-plane depth) of the unit cell, and
s and c are the sine and cosine of /3, respectively. The in-plane macroscopic strains for the unit
cell can be expressed as a function of the displacements in Equation (8) as
2dx1 dx1
x x = = (9)
l R
2d y3 d y3
 yy = = (10)
h 2Rs

Equating the results obtained by Equations (9) and (10) to the expressions for in-plane strains given
by classical 2-D plane elasticity in terms of Youngs modulus and Poissons ratio, the displacements
dx1 and d y3 are found to be
2Rx x
dx1 = (11)
E
2 Rsx x
d y3 = (12)
E
where

plane stress plane strain


E
E = E (13)
(1 2 )

 =  (14)
(1 )

Substituting Equations (11) and (12) into Equation (8), the normal and tangential stiffness in the
unit cell simplify to
1
Kn = E t (15)
3(1  )
(1 3 ) (1 3 )
Kt = K n = E t (16)
(1 +  ) 3(1  2 )

which are independent of the DEM element radius R. Complete details for this process can be found
in Reference [20]. To reiterate, Equations (15) and (16) are the normal and tangential stiffnesses
needed so that a close-packed DEM discretization (Figure 4(b)) gives exact displacement results,

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 387

including the Poisson effect, for a linear elastic material under plane stress/strain conditions
subjected to x-direction uniaxial loading. As will be shown later, Equations (15) and (16) are
also shown to be adequate for irregular discretizations created using the procedure described in
Section 4. For systems with a prescribed element size distribution, a new representative DEM
unit cell should be constructed, and the procedure described above should be repeated in order to
determine a potentially different set of normal and shear stiffnesses.
Other researchers have derived a different expression for the normal stiffness where K n = Et
[15, 21]. It is important to mention that this expression can only be theoretically obtained by
assuming a simple cubic arrangement of DEM elements. The theory provides no basis for the
determination of the shear stiffness, and therefore this parameter is usually arbitrarily chosen as a
fraction of K n [15].
Due to the location and orientation of the contacts in the DEM unit cell, the loading shown
in Figure 4(c) was also considered in order to investigate possible anisotropy in the unit cell.
Surprisingly, the normal and tangential stiffness produced by this exercise agree exactly with
Equations (15) and (16). Assuming the shear stiffness must be non-negative, it is interesting to
note that these equations limit the maximum value of Poissons ratio to  = 1/3 for plane stress and
 = 1/4 for plane strain. Such values correspond to a tangential stiffness of zero. Equations (15)
and (16) were also obtained by Griffiths and Mustoe [22]. However, in their work they used a
different approach in which the strain energy density of three DEM elements in a close-packed
hexagonal arrangement was computed as a function of the normal and tangential stiffnesses and
the in-plane strains of the contact elements.
The DEM unit cell used in this study also proved to be useful to theoretically determine a
time-step bound that ensures stability in explicit time integration, and this topic is discussed in
detail in the following section.

6. STABILITY FOR EXPLICIT TIME INTEGRATION

The subject of DEM numerical stability in explicit time integration is a topic that is usually not
clearly explained in the DEM literature. Explicit integration of the second order equation of motion
by the central difference method is conditionally stable, and the time step t should satisfy


2
t< 1 2
(17)
max
for linear viscous damping, where  is the fraction of critical damping at max , which is the
highest natural frequency of the mesh. Mass proportional damping can be implemented implicitly
if the mass matrix is diagonal, although this improves the time step only marginally since mass
proportional damping decreases at higher frequencies. Nonetheless, if mass proportional damping
is used and is treated implicitly, then Equation (17) applies where  is the fraction of critical
damping at max due to stiffness proportional damping only.

6.1. Estimates of max


For practical problems, max is rarely known and it is customary to use an upper bound for this
frequency for purposes of determining a stable time step size. In finite element analysis, it is usual to
bound max for the mesh by the maximum frequency among all unconstrained individual elements,

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
388 F. A. TAVAREZ AND M. E. PLESHA

Kn
DEM cell unit

Figure 5. DEM discretization and unit cell used for computing the maximum frequency.

(max )e [23]. However, in the DEM, the contact springs are massless, so that (max )e = , and
hence the bound is of little practical usefulness, and other methods must be used.

6.1.1. Traditional estimation of max in DEM computations. Most of the literature on time step
selection for DEM discretizations state that the critical time step is given by
2 2
tc =  (18)
max kmax /m min
where kmaxis the largest inter-element spring stiffness, m min is the mass of the smallest element,
and hence kmax /m min is a crude estimate of the highest natural frequency of vibration for the
model. Moreover, this estimate does not explicitly account for damping. For these reasons, a
user-selected parameter  is included in Equation (18), and computational experience shows that
values of  near 0.1 are typically satisfactory to provide for stable computation [6]. However, the
ad hoc scheme given by Equation (18) is far from ideal, for several reasons. For one, too small a
time step requires unnecessarily long execution time, and for problems with strong nonlinearities,
the stable time step size may change throughout a simulation, and furthermore, if instability does
occur, it may be difficult to detect because of the numerous energy-dissipative mechanisms. In the
following subsection, we discuss a new method for determining the stable time step size.

6.1.2. DEM unit cell approach for estimating (max )e . While it is possible to use the Rayleigh
quotient or Gerschgorin bound to accurately bound max for a DEM discretization [19], the
following approach using a unit cell is effective. The idea rests on the fact that for any discretization,
the imposition of constraints raises the frequency spectrum. Thus, we consider the arrangement
of elements shown in Figure 5, where the seven-element cluster of DEM elements is seen to be
identical to the unit cell used earlier in Figure 4(b). In Figure 5, the frequencies of a constrained
seven element cluster can be computed if all d.o.f. in the full model are constrained, except those
associated with the seven elements. Thus, the highest natural frequency of the seven element
cluster, (max )e , will bound the highest natural frequency of the entire, unconstrained DEM
discretization [24].
Mass and stiffness matrices were computed for the DEM unit cell, which consists of 21 d.o.f.
(14 translational d.o.f. and seven rotational d.o.f.). Values for the stiffness and mass matrices can
be found in [20]. The stiffness matrix for the unit cell was computed using a value for K t equal
to K n , which is the maximum value that K t has for values of Poissons ratio 0. Table I lists
the eigenvalues for the DEM unit cell shown in Figure 5. As shown in the table, the maximum

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 389

Table I. Eigenvalues of DEM unit cell shown in Figure 5.


1 2 3 4 5 6 7 8 9 10 11
2.3542 2.3542 4.7085 5.0 5.0 5.0 5.0 7.0 7.0 7.0 7.0

12 13 14 15 16 17 18 19 20 21
7.6458 7.6458 8.0 8.0 10.0 10.0 14.0 14.0 15.2915 16.0


natural frequency is (max )2e = 16.0K n /m, which gives (max )e = 4.0 K n /m, where m is the
mass of each DEM element and K n is the normal contact stiffness. Substituting this frequency
into Equation (17), we obtain an upper bound estimate for the critical time step tcrit given by


2
tcrit = 2 m/K n
1
1  (19)

It is important to mention that our critical time step bound is a function of the unit cell
configuration used to discretize the medium. For instance, a DEM unit cell having a simple cubic
assembly will produce a different bound for the critical time step than a close-packed assembly.
Note however that a close packed assembly involves the largest number of contacts possible for an
arrangement of elements (assuming constant element diameter), and hence the highest frequencies
are expected for this situation, and Equation (19) should thus apply to other arrangements of DEM
elements. For assemblies with a distribution of particle sizes (hence with varying mass and contact
stiffness values), a DEM element could potentially have a larger number of contacts, and hence
the critical time step should be computed using Equation (18), where an estimate of max would
need to be obtained by other methods.
Related to our approach here is the work of OSullivan and Bray [25] for determining a critical
time step size. In their approach, they consider one contact between two DEM elements, and use
only the normal and tangential direction springs in this contact to determine a stiffness matrix with
of six d.o.f. Because the contact springs have no mass associated with them, simple assumptions
were made to assign portions of the DEM element masses to the d.o.f. based on the number of
contacts for eachDEM element. For a close packed assembly, they report a critical time step
of tcrit 0.408 m/K n (with no damping), which is lower than that given by Equation (19).
However, it is unlikely that this estimate is an upper bound.

7. ELASTIC RESPONSE OF A CANTILEVER BEAM

To test the accuracy of the expressions for contact stiffnesses obtained in Section 5, a cantilever
beam subjected to a slow sinusoidal tip load with frequency  was considered. The accuracy of the
DEM simulations for this problem was studied by comparing the DEM results with simulations
performed using finite element analysis (FEA) with the program ANSYS. In the FEA simulations,
the cantilever beam was discretized using 1440 four-node plane quadrilateral elements.
Figure 6 shows three DEM discretizations for the cantilever beam using close-packed assemblies
with 81, 729, and 6561 elements, respectively, and for each refinement, the element radius was

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
390 F. A. TAVAREZ AND M. E. PLESHA

F = F0 sin t

F = F0 sin t

F = F0 sin t

Figure 6. DEM models of cantilever beam using a close-packed assembly.

0.6 0.6
81 elements 6561 elements
729 elements
0.4 6561 elements 0.4 FEA solution
tip displacement, m

tip displacement, m

0.2 0.2

0 0

-0.2 -0.2

-0.4 -0.4

-0.6 -0.6
0 0.005 0.01 0.015 0.02 0.025 0 0.005 0.01 0.015 0.02 0.025
(a) time, sec (b) time, sec

Figure 7. DEM tip displacement results using a close-packed assembly and comparison with FEA solution.

reduced by a factor of 3. All DEM simulations were performed in plane stress, using the values
for normal and tangential spring stiffness given by Equations (15) and (16) with E = 1.2 GPa
and  = 0.3. The forcing frequency was chosen to be 0.2 times the value of the fundamental
natural frequency of the cantilever beam, and mass proportional damping with a fraction of critical
damping of  = 10% was used at this frequency. Figure 7(a) shows the tip displacement as a
function of time for the three models. Figure 7(b) shows the tip displacement obtained by the
finest mesh along with the FEA solution. As shown in Figure 7(b), the agreement between the
DEM solution and the FEA solution is excellent, and results also appear to converge to a definite
value in the limit of mesh refinement. Applying Richardson extrapolation [19] to the results of

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 391

F = F0 sin t

F = F0 sin t

F = F0 sin t

Figure 8. DEM models of cantilever beam using an irregular element assembly.

these three DEM models provides an estimate of max = 0.4754 m as the steady-state maximum
amplitude of vibration that the DEM models will converge to. The finite element solution provides
FEM = 0.4749 m, which is within 0.1% of the DEM extrapolated solution.
Figure 8 shows three DEM discretizations for the cantilever beam using an irregular assembly
of DEM elements, consisting of 600, 1200 and 2400 elements, respectively. The models were
constructed using the procedure described inSection 4, and for each model refinement, the DEM
element radius was reduced by a factor of 2. For these simulations, the forcing frequency was
chosen to be 0.2 times the value of the fundamental natural frequency of the beam, and mass
proportional damping with a fraction of critical damping of  = 10% was used at this frequency.
Values for normal and tangential spring stiffness were obtained from Equations (15) and (16) using
E = 1.2 GPa and  = 0.3. Figure 9(a) shows the tip displacement as a function of time for the
three DEM models. Figure 9(b) shows the tip displacement obtained by the finest mesh along with
the FEA solution. Even though the results appear to show convergence, it is interesting to note
that contrary to the case using a close-packed arrangement, these models become more flexible as
the mesh is refined. Moreover, results appear to converge to a solution that is considerably lower
(about 6.5%, using Richardson extrapolation) than the FEA solution.
One possible explanation for why the models for the close-packed arrangement become stiffer
as the DEM mesh is refined could be that for each refinement, there are contact springs that are
closer to the top and bottom surfaces of the beam, increasing the flexural stiffness of the model.
This appears to have a greater effect on the behaviour of the model than adding more d.o.f., which
usually results in making a model more flexible. On the other hand, the addition of d.o.f. appears to
have a greater effect on the behaviour of the model when using an irregular element arrangement,
resulting in a numerical solution that becomes softer as the mesh is refined.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
392 F. A. TAVAREZ AND M. E. PLESHA

0.08 0.08

600 elements 2400 elements


0.06 1200 elements 0.06
FEA solution
2400 elements
0.04 0.04

tip displacement, m
tip displacement, m

0.02 0.02

0 0

-0.02 -0.02

-0.04 -0.04

-0.06 -0.06

-0.08 -0.08
0 0.001 0.002 0.003 0.004 0.005 0.006 0 0.001 0.002 0.003 0.004 0.005 0.006
(a) time, sec (b) time, sec

Figure 9. DEM tip displacement results using an irregular element assembly


and comparison with FEA solution.

Table II. Summary of convergence results for the element arrangements used.
Number of elements used
DEM element
arrangement Mesh 1 Mesh 2 Mesh 3  p

Close-packed 81 729 7290 3 1.52


Irregular 600 1200 2400 2 3.00

7.1. Rate of convergence


Using the results of three meshes, the rate of convergence (if uniform) can be estimated using [19]


1
2 1
p = log (20)

2
3 log 
where
i are, in this case, the values of the maximum steady-state displacement for the three
different discretizations, and  is the element size reduction factor for each mesh refinement,
which must be the same for each refinement. As discussed in [19], Equation (20) is accurate only
under fairly strict requirements, including all meshes must be sufficiently fine to yield a constant
rate of convergence. Table II shows the rate of convergence computed for the two different
DEM element packing arrangements used for the cantilever beam. The irregular element-packing
arrangement shows a surprisingly fast rate of convergence with a value of p = 3. However, this
value may be artificially high because the 600 element mesh is too coarse. The rate of convergence
for the close-packed arrangement was p = 1.52, which is a more reasonable value. While these
results are the first of their kind that we are aware of, considerable additional work is still required
to adequately understand the convergence properties of DEM simulations of elastic response
in solids.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 393

8. GENERAL CRITERIA FOR DEM CONTACT FAILURE

Most of the work reported in the literature for modelling fracture in DEM discretizations use a
simple tensile force limit for contact failure [14, 21, 2630]. That is, if the tensile force in a certain
contact exceeds a limit value, the contact breaks and can no longer support tensile forces. This force
limit is usually determined by calibrating DEM simulation results to available experimental data
from tensile and compressive tests for a particular material. Several researchers have imposed limits
for both the tensile and shear strengths in contacts [12, 13, 15]. However, since force transmission
through a DEM medium can only occur via the inter-element contacts, the number of contacts
within a DEM medium will most likely affect the fracture behaviour. Therefore, this contact force
limit most likely depends on element size.
Several researchers have used the maximum normal stress failure model to develop a simple
failure criterion for DEM models that is element size dependent, as follows. Considering a DEM
model consisting of two elements, as shown in Figure 10, failure occurs when the contact stress
reaches the tensile strength of the material [30]. Hence, the tensile force required to break the
contact is given by

f ult = ult (2R)t (21)

where R is the radius of the DEM elements, t is the thickness (out-of-plane depth), and ult
is the tensile strength of the material. Despite the linear dependence on R in Equation (21),
if one considers the contact springs immediately in front of a crack tip, it is clear that use of
Equation (21) will always under-predict the far-field stress that would yield crack growth in a
linear elastic material for sufficiently refined discretizations. In the remainder of this paper, we
develop failure criteria that are objective to mesh refinement, using criteria based on a materials
fracture toughness, K I c , and critical energy release rate, G f .

Kn

t
2R

Figure 10. Tensile strength criterion for failure.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
394 F. A. TAVAREZ AND M. E. PLESHA



+


=

Figure 11. Crack tip stress field for pure Mode I cracking.

The results to be described here begin with the original work of Potyondy and Cundall [15]. In
their work, they postulate a breaking strength between the bonds of contacting DEM elements, and
then address the issue of determining the fracture toughness of the material that is being modelled
by the DEM discretization. The approach described herein differs from theirs in that we view the
fracture toughness as a fundamental material parameter, and then attempt to use theory to develop
expressions for the behaviour of bonds between contacting DEM elements. A main feature of such
an approach is that it promises to provide convergence. That is, it provides a precise specification
of how the contact behaviour between DEM elements must change as a function of element size so
that convergence to the exact solution is achieved as the DEM element size becomes smaller during
model refinement. To model the progressive failure of a solid due to crack growth, a local rupture
criterion is applied to the bonded contacts between interacting elements. The plane stress/strain
near crack tip stress field for pure Mode I loading in a homogeneous, isotropic, linear elastic
material is given by [31]

KI 
(n)
i j (r, ) = Fi j ( ) + An r n/2 Fi j ( ) (22)
2r n=0

Since crack growth is controlled by the stresses near the crack tip, i.e. at small r , the first term in
Equation (22) gives the highest contribution to the stress field at the crack tip. Therefore, keeping
only the first term in Equation (22), the well-known expression for the stresses close to the crack
tip, shown in Figure 11, is obtained. Restricting attention to = 0, the stress  is then

KI
 = (23)
2r

The force required to advance the crack by breaking the contact at the crack tip of a DEM model is
determined by integrating the stress field over the DEM element diameter and thickness as shown
in Figure 11. Therefore,
 t  2R  t  2R
KI
f=  dr dt = r 1/2 dr dt (24)
0 0 0 0 2

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 395

Figure 12. Mixed mode fracture in DEM contacts for a close-packed assembly.


R
f = 2K I t (25)


where f is defined in Figure 10. Assuming linear elastic fracture mechanics (LEFM) conditions,
cracking will occur when the value of K I reaches the materials fracture toughness K I c . Therefore,
the required force to break the contact is given by

R
f ult = 2K I c t (26)


As shown in Equation (26), material failure parameters in this approach are dependent on
DEM element size and the fracture toughness of the material. As described above, this result is
significant because it prescribes exactly how the DEM inter-element strength must change as a
function of element size so that convergence to the LEFM behaviour is obtained as element radius
R vanishes in the limit of mesh refinement. However, it is important to mention that this result was
obtained for a DEM model consisting of a simple cubic element arrangement, and considers only
Mode I cracking, where there is no force being carried by the tangential spring at the crack tip
contact. Since in a two-dimensional simulation both Modes I and II can be present simultaneously,
especially for a close-packed discretization, a fracture criterion should be developed to account for
this type of loading where both the normal and tangential springs at the crack tip contact are being
deformed. For instance, Figure 12 shows a DEM model of a plate with an edge crack subjected
to a uniaxial stress. Even though macroscopically this behaviour would be pure Mode I, since the
far-field stress is perpendicular to the crack plane, both the normal and tangential springs at the
crack tip are neither parallel nor perpendicular to the crack plane. As such, both springs contribute
to resisting the crack advance.
Another important issue to consider is that failure according to Equation (26) assumes that
fracture mechanics controls failure over other phenomena such as the yield strength or maximum
tensile strength. Due to the nature of the DEM medium that is proposed in this work (irregular ele-
ment arrangement with voids), clearly there should be a correlation between the fracture toughness
of the material as a criterion for fracture, and the materials ultimate tensile strength. Figure 13
shows a general DEM model consisting of an irregular assembly of elements of different size. For
such a model, if pre-existing flaws of width a0 populate the medium, the fracture stress should

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
396 F. A. TAVAREZ AND M. E. PLESHA

Figure 13. Irregular assembly of DEM elements of different size.

follow from the fracture toughness K I c of the material as


KIc
ult = (27)
a0

In other words, there is a transition crack length given by


2
1 KIc
at = (28)
 ult
below which there will be little or no strength reduction due to the initial flaws in the material [32].
In a DEM model with no initial flaws, where the fracture toughness is a constant material property,
the failure criterion given by Equation (26), which is based on the materials fracture toughness,
will be inappropriate. For this case, the crack length will be given by the lattice contact spacing,
and therefore the fracture toughness will be element size dependent. Since there are no guarantees
that, in the limit of mesh refinement, the models developed in this work will have flaws greater in
size than that given by Equation (28), an alternate criterion should be developed in order to account
for a condition in which failure will be dictated by the ultimate tensile strength of the material. In
what follows, we develop a contact failure criterion that incorporates both the materials tensile
strength ult and fracture toughness K I c .
The energy release rate in a homogeneous, isotropic linear elastic material is determined by
calculating the total work balance per unit thickness for extending a crack a distance a [31].
Therefore,
 a   f
1
G=  yy d dx (29)
a 0 0

where  f is the crack opening displacement at position x behind the crack tip after crack extension
a, and  yy is a function of both position x and the crack opening displacement  behind the

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 397

crack tip. By solving Equation (29) and taking the limit as a 0, we obtain the well-known
relationship between energy release rate and stress intensity factor K I
K I2
G= (30)
E
where E is given by Equation (13).
In the present work, it is assumed that the energy required to fracture an isotropic material is
independent of the mode of fracture. Therefore, the fracture energy G f or critical energy release
rate is then given by Equation (30) when the value of K I reaches the materials fracture toughness
K I c . That is,
K I2c
Gf = (31)
E
For application in the DEM, a in Equation (29) is replaced by the element diameter and
therefore  yy is averaged by dividing the contact force by the diameter and thickness of the
DEM element. Moreover, the crack opening displacement is also an average computation over the
diameter of the element and consists of the stretch of the normal spring in the crack tip contact.
Therefore, in the present approach, the critical energy release rate is obtained by computing the
area under the curve of  yy vs  at contact failure.
Returning to the DEM unit cell described earlier and shown in Figures 4 and 5, we can obtain
both the normal and tangential contact forces as a function of the applied uniaxial tensile stress
supported by the unit cell. When the applied stress approaches the materials tensile strength ult ,
using the unit cell results for normal and tangential contact stiffness, these forces are


(Rt)ult 
f ncrit = 3 (32)
2(1  ) 3
(Rt)ult
f tcrit = (1 3 ) (33)
2(1  )

where  is given by Equation (14). The derivation of Equations (32) and (33) can be found
in Reference [20]. These critical forces for a close-packed assembly are analogous to the cri-
terion leading to Equation (21), where a simple cubic arrangement is implied. For a flaw-free
medium, contact normal and tangential forces will reach the critical forces given by Equations (32)
and (33) simultaneously and then the model will reach the ultimate tensile strength of the material.
Regardless of the materials fracture toughness, the contact forces should not exceed those given
by Equations (32) and (33). Otherwise, the DEM models tensile strength will be greater than the
materials tensile strength.
When cracks are present, the contact forces at the crack tip reach the critical forces given by
Equations (32) and (33) relatively fast during loading since stresses at a DEM crack tip become
infinite in the limit of mesh refinement. However, it is proposed that complete failure should
not occur before the contact achieves the materials fracture energy G f . In order to satisfy this
condition, the stressdisplacement curve shown in Figure 14 is then adopted for both the normal
and tangential contact at the crack tip. As shown in this figure, the contact springs (both normal and
tangential) behave linearly up to the point where the forces satisfy Equation (32) or Equation (33)

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
398 F. A. TAVAREZ AND M. E. PLESHA

n contact tensile
critical stress
t
contact shear
critical stress
ncrit
tcrit
Gf

n t
ncrit max
n
tcrit max
t

(a) (b)

Figure 14. DEM contact stressdisplacement behaviour and fracture energy.

for the normal and tangential springs, respectively. Once these forces are reached, the behaviour
is proposed to be plastic until the area under the stressdisplacement curve for the normal spring,
shown in Figure 14(a), reaches the fracture energy of the material, given by Equation (31). It
should be noted that the area under the curve in Figure 14(a) is the minimum energy lost, as
energy may also be stored in the tangential direction of the contact. However, this approach proved
to be accurate enough since the critical energy release rate is computed considering pure Mode I
loading only.
The maximum normal displacement nmax is obtained by computing the area under the curve
in Figure 14(a) and equating this to the materials fracture energy G f . Therefore, the maximum
normal displacement is given by

K I2c crit
n
nmax =
+ (34)
crit
n E 2
where

f ncrit ult 
crit
n = = 3 (35)
2Rt 4(1  ) 3

f ncrit Rult
crit
n = = (3  ) (36)
Kn 2E

Equation (34) gives a displacement-based failure criterion, where the maximum displacement is a
function of the materials tensile strength, elastic modulus, Poissons ratio, and fracture toughness
(each of which could be specified independently). Since contact stresses are computed using the
radius of the elements in contact, the failure criterion is also element size dependent. It should
be noted that if unloading occurs in the contact after it reaches the plastic range, the slope of the
unloading forcedisplacement curve will be equal to its initial slope (K n and K t for the normal
and tangential stiffnesses, respectively).
In order to test the failure criterion given in Equation (34), three DEM models representing a
plate with an edge crack subjected to pure Mode I loading were developed. The models, shown in

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 399

radius = 7.812500 mm
2.5 radius = 3.906250 mm
radius = 1.953125 mm

stress, MPa
1.5

0.5

-0.5
0 1 2 3 4 5 6
time, ms

Figure 15. Convergence for crack growth in a DEM model of a cracked plate subjected to
pure Mode I loading; hollow data points in the plot represent the value of far-field stress
that cause unstable crack growth.

Figure 15, consist of 1172, 4713, and 18 899 elements, respectively, arranged in a close-packed as-
sembly. The material modelled by these DEM meshes has elastic modulus E = 29.7 GPa, Poissons
ratio  = 0.3, ultimate tensile strength ult = 3.44 MPa, and fracture toughness K I c = 1.0 MPa m
(i.e. concrete). The edge crack (shown by the solid line at the centre of each model) was cre-
ated by giving contacts along the crack the standard behaviour described by the frictional contact
mechanism shown in Figure 1. All simulations were performed in plane strain, applying a uniform
stress to the top and bottom boundaries, which was slowly increased until failure in the model.
The results of the DEM simulations are also shown in Figure 15, where the hollow data points
represent values of the far-field stress that cause the crack to propagate unstably across the model.
An exact solution for this problem (finite width plate with an edge crack with pure Mode
I loading) does not exist due to the difficulty of finding a solution that satisfies the traction-
free boundary conditions on the free edges. However, several numerical techniques have been
developed in order to obtain very accurate approximate solutions, and a commonly used solution is
given by [33]
KIc
 F = [1.12 0.23(a/c) + 10.6(a/c)2 21.7(a/c)3 + 30.4(a/c)4 ]1 (37)
a
where  F is the value of the far-field stress that causes the crack to propagate unstably across the
plate, a is the crack length (a = 0.03125 m) and c is the width of the plate (a/c = 1/8 for all the
models of Figure 15). Equation (37) was obtained by a least-squares polynomial fit to data obtained
by the Boundary Collocation Method, which is accurate to within 0.5% for a/c0.6. Using the
parameters cited earlier, this equation provides  F = 2.61 MPa. An extrapolated solution based on
the results of the three DEM models, assuming regular mesh refinements, can be computed by

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
400 F. A. TAVAREZ AND M. E. PLESHA

n contact tensile
critical stess

ncrit

Gf
n
ncrit max
n

Figure 16. Alternate DEM contact normal stressdisplacement behaviour.

Richardson extrapolation [19] using


1
3
22

e = (38)

1 2
2 +
3

where in this case,


i represents the value of the far-field stress that causes the crack to propagate
unstably for model i, and
e is the value to be extrapolated. Applying Equation (38) to the results
of the three DEM models in Figure 15 provides an estimate of ult = 2.595 MPa as the value of
the far-field stress that the DEM models will converge to. This value is within 0.5% of the value
provided by Equation (37), which as cited earlier could be in error by as much as 0.5%. Therefore,
the proposed criterion for contact failure appears very promising.
Applying Equation (20) to the results of the three models, a convergence rate p = 1.8 (almost
quadratic) was obtained, which is a reasonable value and is encouraging. The models shown in
Figure 15 were also analysed with no initial crack, and all three models failed at essentially the
same maximum stress of  = 3.44 MPa (independent of model discretization), which agrees with
the value of the ultimate tensile strength of the material ult . Therefore, the proposed method
provides excellent results regardless of the type of failure (fracture toughness criterion or ultimate
tensile strength criterion).
Rather than the plastic behaviour proposed in this work, as shown in Figure 14, several
researchers [3436] have proposed the DEM contact stressdisplacement behaviour shown in
Figure 16, where the contact normal stress decreases linearly to zero after reaching a critical
tensile stress. Analogous to the approach presented in this work, the slope of the curve in the
softening stage is determined such that at failure, the area under the contact stressdisplacement
curve is equal to the materials fracture energy. However, when this contact behaviour was used
in the DEM models shown in Figure 15 with the contact parameters developed in this work
(i.e. contact stiffness K n , and critical tensile stress crit
n , and fracture energy G f ), the DEM results
under-predicted the far-field stress that causes the crack to propagate unstably. That is, even though
the approach could be used in some cases to model material softening phenomena, it does not
appear to guarantee that the DEM model results will converge to the exact solution (assuming
LEFM conditions) in the limit of mesh refinement.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 401

9. DEM FOR MODELLING PENETRATION

To illustrate the application of the DEM method and to demonstrate its utility for modelling chal-
lenging problems, we will consider the example of projectile penetration of a concrete structural
member. While problems of this type have been successfully analysed using FEA (e.g. [37, 38]),
there are numerous difficulties confronting such analyses, including the need for continuum con-
stitutive models which are not well established for many materials of engineering importance, over
the range of conditions and states pertinent to extreme loading.
Figure 17 shows a projectile penetrating a concrete beam at several stages of a DEM sim-
ulation. The concrete beam consists of 12 000 DEM elements and was constructed using the
procedure described in Section 4. The projectile is steel and consists of 351 DEM elements ar-
ranged in a close-packed assembly. While it is possible to allow for damage (e.g. erosion) of
the projectile using the methods discussed in this paper, in this example the projectile undergoes
no damage. As shown in Figure 17, nonlinear phenomena including widespread cracking and
the progressive gross failure and fragmentation of the solid beam into particulates is simulated
automatically.
It is important to mention that projectile impact can generate very high loading rates which can
produce strain rate effects including increased tensile strength, compressive strength, and fracture
energy, in addition to other modifications of material behaviour. Some of these effects have been
included in the results shown in Figure 17. However, because a detailed discussion of these strain
rate enhancements are not central to the main theme of this paper, we refer to Reference [20]
for details.

Stage 1 (t = 0 ms) Stage 2 (t = 1.0 ms)

Stage 3 (t = 2.67 ms) Stage 4 (t = 6.0 ms)

Figure 17. DEM simulation of projectile penetration into concrete beam.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
402 F. A. TAVAREZ AND M. E. PLESHA

10. SUMMARY AND CONCLUSIONS

A DEM was developed as a general computer technique for unified modelling of the mechanical
behaviour of solid and particulate materials, including the transition from solid phase to particulate
phase. Megaclustering was used to model solids, and the method by which a problem domain is
constructed proved to be effective for introducing irregularity in the model, with no anisotropies
or preferential directions for fracture. An energy-based criterion was developed for contact failure
and this criterion was shown to be more appropriate than a criterion based purely in a materials
ultimate tensile strength and fracture toughness. For purposes of selecting a time step size for
integrating the equations of motion by the central difference method, a method was developed
to estimate the maximum frequency of vibration in a DEM model, which gives a more accurate
prediction than the ad hoc procedure that has historically been used.
Using a DEM model of a unit cell of homogeneous material, a theoretical method was developed
to determine the DEM contact normal and tangential stiffnesses as functions of known material
properties so that convergence is obtained with model refinement and the correct elastic behaviour
is produced. To summarize, the normal and tangential stiffnesses are given by

1
Kn = E t (39)
3(1  )
(1 3 ) (1 3 )
Kt =
Kn = E t (40)
(1 +  )
3(1  )2

Regarding contact failure for close-packed and irregular DEM assemblies, a novel DEM contact
behaviour was developed that respects failure governed by the materials fracture toughness and
the materials ultimate tensile strength. The contact stressdisplacement curve is proposed to be
linear until the contact normal and shear stresses reach critical values given by


f crit (Rt)ult 
crit = n where f ncrit = 3 (41)
n
2Rt 2(1  ) 3

f tcrit (Rt)ult
crit
t = where f tcrit = (1 3 ) (42)
2Rt 2(1  )

After the contact forces reach the values given by Equations (41) and (42), the contact behaviour is
proposed to be plastic until the area under the stressdisplacement curve for the normal direction
spring reaches the materials fracture energy, given by

K I2c
Gf = (43)
E
and the maximum normal displacement is then given by

K I2c crit
n f ncrit
nmax =
+ where crit
n = (44)
crit
n E 2 Kn

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
ELEMENT METHOD FOR MODELLING SOLID AND PARTICULATE MATERIALS 403

Even though the simulations showed very good agreement with theoretical results for both
elastic behaviour and fracture, the rate of convergence was not clear, and it appears to depend
on the element arrangement. While these results are the first of their kind that we are aware of,
considerable additional work is still required to adequately understand the convergence properties
of DEM models of solids. It should also be noted that all expressions and criteria presented in
this paper are for two-dimensional DEM models. However, the theory used could in principle be
applied to develop analogous expressions and criteria for three-dimensional systems.

ACKNOWLEDGEMENTS
The financial support of the National Highway Institute in the form of a Dwight D. Eisenhower Trans-
portation Fellowship for Federico Tavarez, and the Computational Mathematics Program of the U.S. Air
Force Office of Scientific Research (AFOSR), Grant No. F49620-03-0216, is gratefully acknowledged.
We are also grateful for the valuable assistance and collaboration of Dr Richard P. Jensen, Sandia National
Laboratory, and numerous conversations with Dr David Potyondy, Itasca Consulting Group, Inc., and Prof.
Walter Drugan, Department of Engineering Physics, University of Wisconsin-Madison.

REFERENCES
1. Cundall PA, Strack ODL. A discrete element model for granular assemblies. Geotechnique 1979; 29(1):4765.
2. Ghaboussi J, Barbosa R. Three-dimensional discrete element method for granular materials. International Journal
for Numerical and Analytical Methods in Geomechanics 1990; 14:451472.
3. Dobri R, Ng T. Discrete modelling of stress-strain behaviour of granular media at small and large strains.
Engineering Computations 1992; 9:129143.
4. Morgan JK, Boettcher MS. Numerical simulations of granular shear zones using the distinct element method
1. Shear zone kinematics and the micromechanics of localization. Journal of Geophysical Research 1999;
104(B2):27032719.
5. Cleary P. Modeling comminution devices using DEM. International Journal for Numerical Methods in
Geomechanics 2001; 25:83105.
6. Jensen RP, Bosscher PJ, Plesha ME, Edil TB. DEM simulation of granular mediastructure interface: effects of
surface roughness and particle shape. International Journal for Numerical and Analytical Methods in Geomechanics
1999; 23:531547.
7. Jensen RP, Edil TB, Bosscher PJ, Plesha ME, Kahla NB. Effect of particle shape on interface behaviour of
DEM-simulated granular materials. The International Journal of Geomechanics 2001; 1(1):119.
8. Thomas PA, Bray JD. Capturing nonspherical shape of granular materials with disk clusters. Journal of
Geotechnical and Geoenvironmental Engineering 1999; 125(3):169178.
9. Boulon M, Nova R. Modeling of soil structure interface behaviour: a comparison between elastoplastic and rate
type laws. Computers and Geotechnics 1990; 9:2146.
10. Jensen RP, Plesha ME, Edil TE, Bosscher PJ, Kahla NB. DEM simulation of particle damage in granular
mediastructure interfaces. The International Journal of Geomechanics 2001; 1(1):2139.
11. Ting JM. A robust algorithm for ellipse-based discrete element modelling of granular materials. Computers and
Geotechnics 1992; 13(3):175186.
12. Potyondy DO, Cundall PA, Lee CA. Modeling rock using bonded assemblies of circular particles. Proceedings
of the 2nd N. American Rock Mechanics Symposium, Montreal, 1996; 19371944.
13. Magnier SA, Donze FV. Numerical simulations of impacts using a discrete element method. Mechanics of
Cohesive-Frictional Materials 1998; 3:257276.
14. Donze FV, Bouchez J, Magnier SA. Modeling fractures in rock blasting. International Journal of Rock Mechanics
and Mining Sciences 1997; 34(8):11531163.
15. Potyondy DO, Cundall P. A bonded-particle model for rock. International Journal of Rock Mechanics and Mining
Sciences 2004; 41:13291364.
16. Cundall PA, Strack ODL. The distinct element method as a tool for research in granular mediaPart I. NSF
Report Grant ENG76-20711, 1978.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme
404 F. A. TAVAREZ AND M. E. PLESHA

17. Cundall PA, Strack ODL. The distinct element method as a tool for research in granular mediaPart II. NSF
Report Grant ENG76-20711, 1979.
18. Williams JR, OConnor R. A linear complexity intersection algorithm for discrete element simulation of arbitrary
geometries. Engineering Computations 1995; 12(2):185208.
19. Cook RD, Malkus DS, Plesha ME, Witt RJ. Concepts and Applications of Finite Element Analysis (4th edn).
Wiley: New York, 2002.
20. Tavarez FA. Discrete element method for modelling solid and particulate materials. Doctoral Thesis, University
of Wisconsin-Madison, 2005.
21. Masuya H, Kajukawa Y, Nakata Y. Application of the distinct element method to the analysis of concrete
members under impact. Nuclear Engineering and Design 1994; 6(2):283294.
22. Griffiths DV, Mustoe GG. Modelling of elastic continua using a grillage of structural elements based on
discrete element concepts. International Journal for Numerical and Analytical Methods in Engineering 2001;
50:17591775.
23. Belytschko T, Liu WK, Moran B. Nonlinear Finite Elements for Continua and Structures. Wiley: New York,
2000.
24. Fried I. Influence of Poissons ratio on the condition of the finite element stiffness matrix. International Journal
of Solids and Structures 1973; 9:323329.
25. OSullivan C, Bray JD. Selecting a suitable time step for discrete element simulations that use the central
difference time integration scheme. Engineering Computations 2004; 21(2/3/4):278303.
26. Zubelewicz A, Bazant ZP. Interface element modelling of fracture in aggregate composites. Journal of Engineering
Mechanics 1987; 113(11):16191630.
27. Bolander JE, Saito S. Discrete modelling of short-fiber reinforcement in cementitious composites. Advanced
Cement Based Materials 1997; 6:7686.
28. Brara A, Camborde F, Klepaczko JR, Mariotti C. Experimental and numerical study of concrete at high strain
rates in tension. Mechanics of Materials 2001; 33:3345.
29. Mishra BK, Thornton C. Impact breakage of particle agglomerates. International Journal of Minerals Processing
2001; 61:225239.
30. Sawamoto Y, Tsubota H, Kasai Y, Koshika H, Morokawa H. Analytical studies on local damage to reinforced
concrete structures under impact loading by the discrete element method. Nuclear Engineering and Design 1998;
179:157177.
31. Anderson TL. Fracture MechanicsFundamentals and Applications (2nd edn). CRC Press: Boca Raton, 1994.
32. Dowling NE. Mechanical Behavior of MaterialsEngineering Methods for Deformation, Fracture and Fatigue.
Prentice-Hall: Englewood Cliffs, NJ, 1993.
33. Cook RD, Young WC. Advanced Mechanics of Materials (2nd edn). Prentice-Hall: Englewood Cliffs, NJ, 1999.
34. Bazant ZP, Planas J. Fracture and Size Effect in Concrete and other Quasibrittle Materials. CRC Press:
Boca Raton, 1993.
35. Mohammadi S, Forouzan-Sepehr S, Asadollahi A. Contact based delamination and fracture analysis in composites.
Thin-Walled Structures 2002; 40:595609.
36. Davie CT, Bicanic N. Failure criteria for quasi-brittle materials in lattice-type models. Communications in
Numerical Methods in Engineering 2003; 19:703713.
37. Schwer LE, Day J. Computational techniques for penetration of concrete and steel targets by oblique impact of
deformable projectiles. Nuclear Engineering and Design 1991; 125:215238.
38. Yadav S, Repetto EA, Ravichandran G, Ortiz M. A computational study of the influence of thermal softening
on ballistic penetration in metals. International Journal of Impact Engineering 2001; 25:787803.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 70:379404
DOI: 10.1002/nme

Вам также может понравиться