Download as pdf or txt
Download as pdf or txt
You are on page 1of 97

Modelling High Density Powder

Compaction Using Multi-Particle Finite


Element Method

Yu Zhang

A thesis in fulfilment of the requirement for the degree of Master

School of Materials Science and Engineering


Faculty of Science
The University of New South Wales
February 2018
PLEASE TYPE
THE UNIVERSITY OF NEW SOUTH WALES
Thesis/Dissertation Sheet

Surname or Family name: Zhang

First name: Yu Other name/s:

Abbreviation for degree as given in the University calendar: Masters of


Engineering

School: School of Materials Science and Engineering Faculty: Faculty of Science

Title: Modelling High Density Powder Compaction Using Multi-Particle Finite


Element Method

Abstract 350 words maximum: (PLEASE TYPE)

Multi-Particle Finite Element Method (MPFEM) is a sub-particle scale numerical approach to provide an accurate description of particle deformation, thus well
suits for modelling high density compaction in which particles experience large deformation. In this work, a 3D MPFEM model was developed to investigate the
structure and stress of particle compacts at high density compaction.

The model was firstly developed to simulate the compression of single particles and a BCC ordered packing. An elastoplastic model was adopted in the model.
The simulation results such as the stress-strain relation and the variation of packing structure were compared with previous studies to validate the model. The
model was then applied to simulating the die compaction of a random packing with an initial packing state generated using the discrete element method (DEM).
The results showed the relative density increased with the increasing pressure applied on the punch. During unloading, the pressure sharply dropped to zero with
slight decrease in compact density. The forces between particles were analysed, showing they were mainly vertical and increased with increasing density.

The model was further extended to study the softening effect of particles and the mixing of particles of different hardness on compaction. In the particle softening
study, the hardness of the particles varied with temperature. With increasing temperature and decreasing hardness, the hardness affected the porosity of packing
dramatically. Larger hardness led to a higher porosity under the same pressure while force orientation and coordination number showed no much difference. The
results of mixture model showed that the soft particles took most responsibility of deformation and the hard particles deformed little. However, the stress in the
hard particles were larger than the soft particles, indicating that the forces were transmitted in compacts mainly through the hard particles.

Declaration relating to disposition of project thesis/dissertation

I hereby grant to the University of New South Wales or its agents the right to archive and to make available my thesis or dissertation in whole or in part in the
University libraries in all forms of media, now or here after known, subject to the provisions of the Copyright Act 1968. I retain all property rights, such as patent
rights. I also retain the right to use in future works (such as articles or books) all or part of this thesis or dissertation.

I also authorise University Microfilms to use the 350 word abstract of my thesis in Dissertation Abstracts International (this is applicable to doctoral theses only).

…………………………………………………………… ……………………………………..……………… ……….……………………...…….…


Signature Witness Signature Date

The University recognises that there may be exceptional circumstances requiring restrictions on copying or conditions on use. Requests for restriction for a
period of up to 2 years must be made in writing. Requests for a longer period of restriction may be considered in exceptional circumstances and require the
approval of the Dean of Graduate Research.

FOR OFFICE USE ONLY Date of completion of requirements for Award:


COPYRIGHT STATEMENT

‘I hereby grant the University of New South Wales or its agents the right to
archive and to make available my thesis or dissertation in whole or part in the
University libraries in all forms of media, now or here after known, subject to the
provisions of the Copyright Act 1968. I retain all proprietary rights, such as patent
rights. I also retain the right to use in future works (such as articles or books) all
or part of this thesis or dissertation.
I also authorise University Microfilms to use the 350 word abstract of my thesis in
Dissertation Abstract International (this is applicable to doctoral theses only).
I have either used no substantial portions of copyright material in my thesis or I
have obtained permission to use copyright material; where permission has not
been granted I have applied/will apply for a partial restriction of the digital copy of
my thesis or dissertation.'

Signed ……………………………………………...........................

Date ……………………………………………...........................

AUTHENTICITY STATEMENT

‘I certify that the Library deposit digital copy is a direct equivalent of the final
officially approved version of my thesis. No emendation of content has occurred
and if there are any minor variations in formatting, they are the result of the
conversion to digital format.’

Signed ……………………………………………...........................

Date ……………………………………………...........................
ORIGINALITY STATEMENT
‘I hereby declare that this submission is my own work and to the best of my knowledge
it contains no materials previously published or written by another person, or substantial
proportions of material which have been accepted for the award of any other degree or
diploma at UNSW or any other educational institution, except where due
acknowledgement is made in the thesis. Any contribution made to the research by
others, with whom I have worked at UNSW or elsewhere, is explicitly acknowledged in
the thesis. I also declare that the intellectual content of this thesis is the product of my
own work, except to the extent that assistance from others in the project's design and
conception or in style, presentation and linguistic expression is acknowledged.’

Signed ……………………………………………..............

Date ……………………………………………..............

i
Abstract
In powder metallurgy, loose particles are compressed into a compact with high density.
During compaction, particles experience large deformation which should be accounted
for in order to understand the compaction behaviour of particles. In this work, a
numerical approach based on Multi-Particle Finite Element Method (MPFEM) was
developed to simulate high density compaction of particles.

The model was firstly validated by simulating the compression of single particle and a
body centred cubic (BCC) packing. The elasto-plastic force model was adopted in the
model and the parameters were fit from previous results. The simulation results were in
good agreement with previous work, thus confirming the validity of the model. The
independence of the results to mesh size, time speed and strain rate were also tested.
The analysis of the Mises stress inside the particle showed that the stress mainly
occurred near the contact surface. The model was then applied to simulating die
compaction of a random packing. The particles experienced local rearrangement and
deformation during the compaction. The inter-particle overlaps increased with
increasing density. The forces between particles were mostly vertical.

The softening effect of particles at high temperature during compaction was also
investigated. Packing and compaction of wax particles at different temperatures (35℃,
37℃ and 39℃) were simulated. In the packing process, the particles were consolidated
under the gravity. While particle hardness had a significant effect on compact density,
the stress distribution in the three cases were similar. In compaction, different pressures
were applied and the final densities of the compacts were comparable to the
experiments. The analysis of forces, however, showed no much difference for the three
cases. So while particle hardness played an important role in packing density, it had
little effect on the forces between the particles.

The simulation of the mixture of soft and hard particles showed that the soft particles
took the most responsibility for deformation and the hard particles were only slightly
deformed. However, the forces on the hard particles were much larger than the soft
particles. Compared to a system which contained only hard particles, much smaller
pressure was required to achieve the same density. The different rearrangements of the

ii
two packings also led to the difference in force orientation and coordination number
distributions.

iii
Acknowledgement
I would like to express my great gratitude to my supervisor, A/Prof. Runyu Yang of
University of New South Wales who gave me a lot of professional guidance to help me
finish my master degree.

There are many thanks to my colleagues from whom I have received help and care.
Thanks to my friend Chenliang Li, Yaoyu Li for their suggestions on the project. I am
also grateful for the encouragement and support of the staff in the School of Materials
Science and Engineering in UNSW during the work.

Last but not least, I am thankful to my parents and my girlfriend for their unconditional
support and love.

iv
Table of Contents
Abstract .............................................................................................................................ii

Acknowledgement ........................................................................................................... iv

List of Figures .................................................................................................................vii

List of Tables ................................................................................................................... xi

Chapter 1. Introduction ..................................................................................................... 1

Chapter 2. Literature Review ............................................................................................ 4

2.1 MPFEM fundamentals and governing equations .................................................... 4

2.1.1 Fundamentals of MPFEM ................................................................................ 4

2.1.2. Governing Equations ...................................................................................... 4

2.2 2D MPFEM Modelling ........................................................................................... 5

2.2.1 Single particle compaction ............................................................................... 5

2.2.2 Multiple particle compaction ........................................................................... 8

2.3 3D MPFEM modelling ......................................................................................... 14

2.3.1 Modelling of single particle compression ...................................................... 14

2.3.2 Modelling of multiple particle compression .................................................. 17

2.4 Summary ............................................................................................................... 20

Chapter 3 MPFEM model development and validation ................................................. 22

3.1 Introduction ........................................................................................................... 22

3.2 Model development .............................................................................................. 23

3.2.1 Constitutive model in MPFEM ...................................................................... 23

3.2.2 Simulation condition ...................................................................................... 24

3.3 Results and discussion .......................................................................................... 27

3.3.1 Model validation ............................................................................................ 27

3.3.2 Compaction of random packing ..................................................................... 36

3.4 Conclusion ............................................................................................................ 42

Chapter 4 Effects of particle hardness on compaction .................................................... 43

v
4.1 Introduction ........................................................................................................... 43

4.2 MPFEM model and simulation conditions ........................................................... 44

4.2.1 Governing equations ...................................................................................... 44

4.2.2 Simulation conditions .................................................................................... 45

4.3 Results and discussion .......................................................................................... 48

4.3.1 Effect of particle softening ............................................................................. 48

4.3.2 Effect of particle mixture ............................................................................... 58

4.4 Conclusion ............................................................................................................ 63

Chapter 5 Conclusions and future work .......................................................................... 65

References ....................................................................................................................... 67

Appendix A.Script to generate a random packing in Abaqus ...................................... 71

Appendix B. Script to of probability density calculation................................................ 80

vi
List of Figures

Fig. 2.1 Three different meshes: (a) coarse mesh, (b) advanced course mesh, and (c) fine
mesh. Figure reproduced from Ref. (Zhang 2009). ············································ 6

Fig. 2.2 MPFEM mesh convergence results (Sinha 2011). ·································· 6

Fig. 2.3 (a) Mises stress, (b) hydrostatic stress (c) tensile stress from elastic simulation
and (d) Mises stress from elastic-plastic model. Figure reproduced from Ref. (Procopio
et al. 2003). ·························································································· 8

Fig. 2.4 Predicted mean stress and relative density in several cases (Procopio 2006). · 10

Fig. 2.5 Relations between pressure and relative density from numerical simulation and
physical fitting (Zhang et al. 2015). ···························································· 11

Fig. 2.6 Relative density of each powder in different punch speed (Kyung-Hun et al.
2009). ······························································································ 12

Fig. 2.7 Behaviour and effective stress distributions of powders at punch speeds of (a) 5
mm/min and (b) 60 mm/min. Figure reproduced from Ref. (Kyung-Hun et al. 2009). 13

Fig. 2.8 Relationship between particle number and normalized axial stress (■) and the
computing time (□) in MPFEM (Procopio and Zavaliangos 2005). ······················· 14

Fig. 2.9 Comparison of the numerical results with two friction coefficients and
experimental results. (Chen et al. 2007). ····················································· 15

Fig. 2.10 Mesh sensitivity tests of (a) one particle and (b) two particles (Frenning 2008).
······································································································· 16

Fig. 2.11 Comparison of the force-displacement behaviour of hydrostatic compaction,


die compaction and simple compression (Harthong et al. 2009). ·························· 17

Fig. 2.12 Comparison of simulated and measured force-displacement. The insets


showed the initial and final states of the compact. ··········································· 18

Fig. 2.13 Comparison between the results of DEM with new contact law, the MPFEM
results and the Storakers's model result (a) lead parameters; (b) MPFEM result with
copper parameters and experiments (Jerier et al. 2011). ···································· 20

Fig. 3.1 Compactions of powders: (a), single particle compression; (b), BCC ordered
packing; and (c), random packing. ····························································· 24
vii
Fig. 3.2 Stress-strain relationship for lead spheres (Fitting using Eq. 4.2 with K= 15.5,
n=0.35) experiment result from (Chen et al. 2007). ········································· 26

Fig. 3.3 Comparison of the (a) force-displacement; and (b) stress-strain relations for
compaction of a particle (Chen et al. 2007). ·················································· 28

Fig. 3.4 Mesh design with element size of (a) 1.2mm; (b) 0.8mm; (c) 0.4mm. ········· 28

Fig. 3.5 Force-displacement curves under different conditions: (a) mesh sizes; (b) time
steps; and (c) punch speeds. ····································································· 30

Fig. 3.6 Von Mises distribution (a) sphere compressed in simulation; (b) sphere
compressed in (Chen et al. 2007). ······························································ 31

Fig. 3.7 Mises stress and pressure of one particle compressing at different displacements
Mises stress: (a) 1mm; (b) 6mm; (c) 10mm. Pressure: (a’) 1mm; (b’) 6mm; (c’) 10mm.
······································································································· 32

Fig. 3.8 Stress linearization on lines across the centre of particle (a) horizontal line; (b)
vertical line. ······················································································· 34

Fig. 3.9 (a) Comparison of the pressing force of simulation and Chen et al. result (Chen
et al. 2007); (b) Pressure-strain results of BCC ordered simulation. ······················ 35

Fig. 3.10 The final state of compaction (a) the current study; (b) study of Chen. Centre
sphere of (a’) 9-particle packing; (b’) simulation results from Chen. Figures b and b’ are
reproduced from Ref. (Chen et al. 2007). ····················································· 36

Fig. 3.11 Pressure as a function of density. ··················································· 37

Fig. 3.12 Stress distributions at different stages of the compaction of the random
packing: (a) density = 0.55; (b) density = 0.7 (c) density = 0.9 (d) density ≈ 0.9. ······ 38

Fig. 3.13 Distribution of overlap and neighbouring particle distance. ···················· 39

Fig. 3.14 Force angle distribution of four stages in simulation of random packing. ···· 40

Fig. 3.15 Curve of coordination number and number of particles. ························ 41

Fig. 3.16 Relationship of overlap and magnitude of force. ································· 42

Fig. 3.17 Distributions of the contact forces at different stages. ··························· 43

Fig. 4.1 Strain-stress curve of different hardness. ··········································· 47

viii
Fig. 4.2 Initial configuration of the particles. Black particles are hard and white particles
are soft. ····························································································· 48

Fig. 4.3 Strain-stress relations of soft and hard particles. ··································· 49

Fig. 4.4 Forces on the bottom wall for the packing of different temperatures (35 ℃, 37 ℃
and 39 ℃). ························································································· 50

Fig. 4.5 Final states at different temperatures after packing under gravity: (a) 35℃; (b)
37℃; (c) 39℃. The colour represents the magnitudes of the normal forces. ············· 50

Fig. 4.6 Distributions of the inter-particle overlap in three packings. ····················· 51

Fig. 4.7 Probability density and force angle in three packings before compression. ··· 52

Fig. 4.8 Distribution in coordination number for three packing before compression. ·· 52

Fig. 4.9 Fitting results of final porosity and pressure on punch for experiments (Zheng
2007) and simulations in different temperature. The line indicates shows the compact
structure at the pressure of 1.2KPa. ···························································· 53

Fig. 4.10 Compact structure at the pressure of 1.2KPa: (a) 35℃; (b) 37℃; (c) 39℃. ·· 55

Fig. 4.11 Probability density function of overlaps in the compacts. ······················ 55

Fig. 4.12 Probability density and force angle in three packings after compression. ···· 56

Fig. 4.13 The distribution in coordination number for three packings after compression.
······································································································· 57

Fig. 4.14 Curves of force and overlap ratio in one particle compression with three
hardness and fitting curves. ····································································· 58

Fig. 4.15 Distribution of normalized contact force between particles in models with
different hardness under same pressure. ······················································· 59

Fig. 4.16 Curve of relative density and upper punch pressure in hardness model
including loading and unloading. ······························································· 60

Fig. 4.17 Compact structure and Mises stress distributions at different stages: (a) stage
A; (b) stage B; (c) stage C; (d) stage D. ······················································· 61

Fig. 4.18 Pressure distribution of four stages in hardness model results. (a)stage A; (b)
stage B; (c) stage C; (d) stage D. ······························································· 62

ix
Fig. 4.19 Distribution of overlap in hardness model. ········································ 62

Fig. 4.20 Force angle distribution of hardness model at four stages. ······················ 63

Fig. 4.21 Coordination number and particle numbers in hardness model. ··············· 64

x
List of Tables

Table 3.1 Parameters used in present work. ·················································· 26

Table 4.1 Parameters of in the particle softening simulations. ····························· 46

Table 4.2 Parameters of hardness model. ····················································· 48

Table 4.3 Fitting parameters of three models. ················································ 57

xi
Chapter 1. Introduction
Powder metallurgy (PM) is a process that uses powders as raw material and sintering of
metal materials, form composite materials and process various types of products
(German 1984). Generally, it contains several steps: powder preparing, powder packing
and compaction and heat treatment. PM is now used in a variety of products and plays
an important role in many industries. The materials being processed include metals and
metal composites material, metal and non-metallic composite materials, metal-ceramic
composite material and so on.

Powder compaction is a significant stage in PM and has a significant effect on the


properties of final parts (German 2005). A typical die powder compaction process
involves three major stages:

• Die Filling: filling amount of loose powders into a die cavity.

• Compaction: the powders are compacted via the loading of punches.

• Unloading and Ejection: the punch moves upwards and the compact is removed.

Many changes are involved in compaction, such as particle deformation, particle


movement, inter-particle bonding and densification. During the densification, the stress
of particles become very complex, and this results in intricate anisotropic density and
stress structure in the final compacts. Many researchers sought to understand the
mechanisms of the compaction process through experimental work or numerical
simulations.

Modelling of powder compaction is an effective way to explore the effects of the


parameters and properties of compaction. In terms of modelling approaches, there are a
few methods, such as Finite Element Method (FEM) (Sokolnikoff and Specht 1956),
Discrete Element Method (DEM) (Aydin et al. 1997). FEM, also known as finite
element analysis (FEA), is a numerical method for solving problems in engineering by
creating a mesh to divide the whole problem into smaller elements. But in powder
compaction, FEM cannot describe the interaction between particles. DEM has been
used by many researchers to simulate powder compaction, it contains several individual
particles and applies contact laws between these particles (Cundall and Strack 1979).
However, both methods cannot provide accurate information with large deformation in

1
powder compaction. When the relative density is large, inter-particle contacts start to
overlap and there is currently no governing equation that can describe the 3-body
interaction. Therefore, new numerical approaches are required to simulate high-density
compaction.

Multi-Particle Finite Element Method (MPFEM) was firstly proposed by Ransing et al.
(2000) to simulate an assembly of particles. This approach discretises individual
particles and the interactions between particles are modelled at the cell level which can
be described by well-established mechanics. MPFEM has been demonstrated to be a
very efficient tool to combine the advantages of FEM and DEM (Zhang et al. 2015). It
can reveal the details of particle deformation and stress in high-density compaction. As
computational cost is a major constraint in MPFEM, the number of particles in powder
compaction simulation using MPFEM is much smaller than in DEM.

MPFEM models provide a great degree of control of properties like:

• behaviour and material properties of particles,

• friction between particles,

• initial microstructure and particle arrangement,

• loading conditions

In this study, particle compression will be carried out in 3D simulations using


ABAQUS software to further study the micro-mechanics and improve the
understanding of the compaction process with multi-particle finite element method.

The goal is to develop an MPFEM model to investigate the structure and stress of
compacts at high-density compaction and to explore mechanical behaviour of particles
in powder compaction.

The specific aims of this work are:

• To develop and validate a 3D MPFEM model for high-density particle compaction;

• To analyse the structural and stress properties of compacts; and


• To investigate the effect of particle hardness on compaction behaviour. including
softening process.

2
The structure of the thesis is as follows:

Chapter 2 reviews previous works on powder compaction using MPFEM including the
fundamentals of MPFEM, compaction of single particles, 2D and 3D compactions.

In Chapter 3, an MPFEM model is developed to simulate the compression of single


particles, compaction of an ordered packing to verify the accuracy of the model. The
model is then extended to compaction of a random packing. The stress and structure of
the compact is analysed.

Chapter 4 presents a study on the effect of particle hardness on compaction. The


softening of particles at high temperature and the mixture of particles of different
hardness are investigated.

Chapter 5 summarises the main the findings of the work and provides some
recommendations for future work.

3
Chapter 2. Literature Review

This chapter reviews the previous work on MPFEM modelling of powder compaction.
The fundamentals of MPFEM will be introduced first, then the applications of MPFEM
to 2D and 3D compactions will be discussed.

2.1 MPFEM fundamentals and governing equations


2.1.1 Fundamentals of MPFEM

In terms of MPFEM, particles are discretised individually using mesh so that the
deformation of particles in powder compaction can be well described. Therefore, not
only the particle-particle pressures can be calculated, but the pressure inside particles
can also be described by means of mesh deformation.

2.1.2. Governing Equations

The deformation of a particle includes both elastic deformation and plastic deformation,
given by.

ε = εel + εpl (2.1)

where ε is the total strain, εel and εpl are the elastic and plastic strains, respectively.

The elastic deformation is described by Hook’s law (Güner, 2015), given by:

σel = Eεel (2.2)

where σel is the elastic stress, E is the Young’s modulus of the material.

A few plastic models have been proposed to describe the plastic deformation of
materials. In the von-Mises model, the stress-strain relation is given by (Güner, 2015):

σY = A(ε0 + εe )m + Bεnr (2.3)

where σY is yield stress, ε0 is initial yield strain, εe is equivalent strain, and εr is


equivalent strain rate. The A, B, m and n are material constants. When the equivalent
strain and stain rate is zero in initial state, the initial yield strain can be calculated by A,
m and Young’s modulus.

Another widely used model is Shima-Oyane model that is suitable for copper
material powder (Shima, 1976). The model is given by:

4
q p
ϕ = (σ )2 + α(1 − D)γ (σ )2 − Dm (2.4)
Y Y

where D is relative density, m, α and γ are parameters calibrated from experiments.

2.2 2D MPFEM Modelling


2.2.1 Single particle compaction

MPFEM modelling single particle compression is a fundamental study that should be


conducted before modelling the compaction of multiple particles. The study on single
particle compression can determine proper model parameters, including mesh size and
time step. The modelling results can be compared with theoretical calculation to verify
the validity of models. The effect of material properties can also be easily studied with
single particle compression. For example, Procopio et al. (2003) studied the effect of
material properties on deformation and the stress distribution in a particle under
compression. Due to the fact that the particle was symmetric, only a quarter of the
sphere was meshed. Their results showed an excellent agreement with experimental
results.

Meshing is an important step in MPFEM modelling and various meshing techniques


have been developed. Since the main deformation occurs at the surface area of particles
in low density compaction, inter-particle elements can be limited to balance the
calculation time. Fig. 2.1 shows three types of mesh and showed the differences of these
meshes in Zhang’s work (Zhang 2009). Fig. 2.1a is a mesh used in many previous
works, such as that of Lee et al. (2009). While the surface elements are smaller than the
internal elements to improve the simulation accuracy, the overall sizes of the elements
are still too large to achieve results comparable to experimental observation. Zhang
(2009) adopted a method to have more elements on the surface of a sphere to improve
the accuracy (Fig.2.1b). Meanwhile the cost of computation was not increased too much
when compared with fine mesh. But this type of mesh was obvious cannot be used for
high density models which are used to simulate systems with a relative density higher
than 0.85 when the overlaps of particles start to interact. The fine mesh in Fig.2.1c was
used in the modelling of particle deformations (Mesarovic and Fleck 2000). Fine mesh
limits the number of particles with longer simulation time.

5
Fig. 2.1 Three different meshes: (a) coarse mesh, (b) advanced course mesh, and (c) fine
mesh. Figure reproduced from Ref. (Zhang 2009).

The influence of mesh size is significant in terms of designing mesh for particles in
powder compaction. To determine what mesh size is suitable, Sinha (Sinha 2011)
studied the relationship between accuracy and the number of elements in mesh. In the
simulations, four different meshes of a sphere with the radius of 10 mm with different
numbers of nodes were compared (76, 420, 840, 1000 surface nodes). Fig. 2.2 shows
the contact force versus displacement curves of different cases. As the elements
increased, the accuracy improved. The results with 76 surface nodes had the most
fluctuated curve. The total trend was nearly the same. He used the 76 surface nodes
mesh in his work in order to balance the efficiency and accuracy.

40
76 surface nodes
35 420 surface nodes
840 surface nodes
30 1000 surface nodes
25
Force (KN/m)

20

15

10

0
0 0.2 0.4 0.6 0.8 1
Displacement (m)

Fig. 2.2 MPFEM mesh convergence results (Sinha 2011).

6
Stress distribution can be well illustrated with one particle model using MPFEM.
Procopio et al. (2003) studied a particle compressed by two rigid planes until the final
displacement of the planes was 10% of the particle diameter. The Mises stress is also
referred to as the equivalent tensile stress and is used to predict the yielding of materials
subject to multiaxial loading conditions, pressure and the principal tensile stress of
models.

Fig. 2.3 indicated that the stress distribution of the FEM simulation results agreed with
descriptions of previous works and theoretical results (Lautenschlager and Harcourt
1970, Aydin et al. 1994). Fig. 2.3a shows the maximum of Mises stress under the
surface of contact area and stress decreases to surface and centre of particles. In Fig.
2.3b, the maximum pressure took place at the surface of the contact area which results
in a deformation of particles. Fig. 2.3c showed tensile stress at transverse direction
reaching the maximum area at the centre of particle and the results were in line with the
prediction of Hertz and Hondros solutions. This stress was responsible for the brittle
fracture of the particle. The stress distribution matches earlier experimental results and
theoretical work. Fig. 2.3d shows the Mises stress of elastic-plastic simulation. It
indicated that the plastic deformation of particle appeared primarily below the contact
surface (the area of maximum Mises stress) and expanded towards the centre since
deformation mainly occurred under the contact surface. The shape and location of
maximum Mises stress areas of the elastic model and the elastic plastic model are
different from the main deformation area in the plastic model which is larger and the
stress is smaller with the same displacement.

7
Fig. 2.3 (a) Mises stress, (b) hydrostatic stress (c) tensile stress from elastic simulation
and (d) Mises stress from elastic-plastic model. Figure reproduced from Ref. (Procopio
et al. 2003).

2.2.2 Multiple particle compaction

To study particle behaviour in powder compaction by considering the interaction of


particles, Ransing et al. (2000) proposed a combined discrete and continuum model.
They made use of two types of coarse mesh to describe two different particles. The
ductile particle was mapped by quadrilateral elements and the brittle was mapped by
triangular elements. The results of this case showed that this model is capable of
simulating the compression of particulate systems. The results also showed the ability
of this method to investigate the interaction of particles. However, the meshes were not
symmetric. This may lead to some numerical errors and they were too coarse to
generate results of enough accuracy.

Procopio (Procopio 2006) compared predicted stress and relative density by MPFEM to

8
test its dependability with FEM and prior models. The findings were shown in Fig. 2.4.

 m  ( z   x) / 2 is the hydrostatic component of the macroscopic stresses in the x


and z-directions  x and  z . y is the yield pressure of the particles. As the relative

density increased, the growth of predicted stress increased. Furthermore, as the relative
density nears 1, the stress tends to be infinity. Since the densification occurred with
particle rearrangement, the multiple particle packing predicted less stress in MPFEM.
The Torre and modified Gurson models which were considered suitable model for large
relative density were now overestimated the mean stress for all the values of the relative
density. Owing to the geometric effects this may be different. The two models treated
the system as a cylinder with a hole at the centre and applied pressure everywhere along
the periphery of the hole. From this figure, the MPFEM was accurate in describing not
only low-density compaction but also high-density compaction. Two material property
combinations were used in the unit cell finite element method of periodic hexagonal
arrangement of discs. The mean stress of two FEM mode firstly increased faster than
MPFEM and as the relative density increased, the increase slowed and get closer to
MPFEM.

Fig. 2.4 Predicted mean stress as a function of the relative density for different cases
(Procopio 2005).

Following the validation of MPFEM, some effects of parameters in powder compaction


can be assessed using this method such as particle deformation change with various
structures or different punch speed. Usually, the powder is densified during powder

9
compaction and the types of densification mainly undergo three steps: particle
rearrangement, elastic deformation and plastic deformation (Kyung-Hun et al. 2009).

Zhang et al. (2015) considered the microscopic behaviour of 2D particle compaction


using MPFEM of four initial packing structures (tetragonal packing, HCP, honey comb
packing, random packing). The packings were produced by DEM simulation except for
the honeycomb packing which was manually generated. The packings were placed in
rigid walls which were fixed with only upper punch moved downward. Fig. 2.5 shows
the four structures and changes of arrangement of particles as well as Mises stress
distribution during compaction. There was only deformation and no relative sliding in
tetragonal packing and HCP (Fig. 2.5), however in random packing and honey comb
packing, the deformation and relative sliding and rolling among particles occurred.
After packing, the shapes of particles in compacts of tetragonal packing and HCP were
rectangles and hexagons. Nonetheless, even honeycomb packing was also an ordered
packing structure. The densification behaviour was different from tetragonal and HCP.
During the early stages of honeycomb packing, the structure was constant, however as
the pressure increased the order was destroyed and rearrangement took place. In this
case, the rise in pressure was slow in comparison to tetragonal and HCP structure. For
random packing compaction, slipping of particles took place at the beginning of
compaction and no obvious deformation of particles was seen. After this stage, the
particles were in a jammed state, the pressure increased rapidly and plastic deformation
became main deformation. At the final stage shown in Fig. 2.5, all the particles had
different shapes due to the different interactions between their neighbouring particles.
Zhang et al. evaluated the force behaviour among four initial packing structures and
came to the conclusion that these four structures lead to different stress distribution and
different force chains in particles that determined the final compact structure and
property. Fig. 2.5 showed the curve of relative density and pressure from simulations
and experiments. Similar trends occurred for each structure. The initial relative density
of four structures were all different. Initially, the pressure increased slowly when the
relative density was low, then at high relative density, the pressures from numerical
simulations tended to be the same.

10
900
Tetragonal
800
HCP
700
Honey-comb structure
600
Disordered structure
Pressure, (MPa)

500 Physically fitted curve


400

300

200

100

0
0.55 0.65 0.75 0.85 0.95
Relative Density

Fig. 2.5 Relations between pressure and relative density from numerical simulation and
physical fitting (Zhang et al. 2015).

Punch speed is a crucial factor that affects the deformation of particles during powder
compaction. Kyung-Hun et al. (Kyung-Hun et al. 2009) studied the effects of various
punch speeds on deformation behaviour and densification of Al particles. The
simulations were carried out for different punch speeds of 5mm/min, 15mm/min,
30mm/min and 60mm/min. The researcher compared the load-stroke curves of particles
of compaction with different punch speeds. Fig. 2.6 shows the results regarding the
relationship between rolling moment, as the relative density and punch speeds (Kyung-
Hun et al. 2009). Higher compaction velocity decreased both the relative density of
final powder compacts and the particle rearrangement. When the punch speed was
higher, the rolling moments and particle movement decreased g, so that the powder
densification was lower. This may be attributable to the fact the high speed gave less
time for the rearrangement of particles. From the figure, it is clear that the decrease of
particle size leads to initial packing density increase. The relative density in
experiments is much lower than simulations, and there is a sudden change at punch
speed is 15mm/min in the experiment results, which may be because the particle
rearrangement is lacking when the punch speed is over 15mm/min.

11
Al 1100 powder, 20µm
0.96 Al 1100 powder, 3µm
Particle diameter, 1.0mm
Particle diameter, 0.5mm
0.92
Relative density

0.88

0.84

0.8
0 10 20 30 40 50 60
Punch speed/(mm•min-1)

Fig. 2.6 Relative density of each powder in different punch speed (Kyung-Hun et al.
2009).

Fig. 2.7 showed the behaviour and stress distribution of powders. The punch speeds
were 5 and 60 mm/min respectively and the particles diameter was 1 mm. As the punch
stroke increased, the powder densification was primarily particle rearrangement and
elastic deformation at first when the punch speed was 5 mm/min. On the other hand,
there was very little particle rearrangement and the main factor of densification is
particle plastic deformation. At higher punch speeds, the period of particle
rearrangement was limited. Particles near boundaries had a higher plastic deformation
and the stress was higher as a result of the friction between the particles and the
boundaries.

12
Fig. 2.7 Behaviour and effective stress distributions of powders at punch speeds of (a) 5
mm/min and (b) 60 mm/min. Figure reproduced from Ref. (Kyung-Hun et al. 2009)

Because of the smaller particle size, the densification of powder that took place was
mainly particle rearrangement in comparison to Fig. 2.7. Following the rearrangement,
the friction increased causing the loads of compaction to increase radically, since the
contact area of the particle increased. The porosity of the small particle model is
noticeably smaller than when the particle diameter is 1 mm under the same punch stroke
(Kyung-Hun et al. 2009).

Moreover, a convergence study of the effects of particle number when evaluating


MPFEM under die compaction is necessary and is presented in Procopio and
Zavaliangos’s work (2005). Fig. 2.8 showed the result that with different numbers of
particle (n= 100, 200, 400, 800, 1600), at final relative density of 0.95, normalized axial
stress and computation time. The number of particles had a significant effect on the

13
computation time. The normalized axial stress decreased with the number of particles
because of the wall effect such that the particles near the wall in the small system (like
n=200) were less and have larger deformations than in the bulk. The results of the
simulation were dramatically affected by the number of particles when the number was
small. When the number was larger than 800. The results showed a small error in
comparison. In other words, the 800 particles simulation was shown to be adequate for
this study.

1.15 1000

1.1

1.05

1 100

CPU Time (hrs)


0.95
Σm/σy

0.9

0.85 10

0.8

0.75

0.7 1
100 400 700 1000 1300 1600
Number of particles

Fig. 2.8 Relationship between particle number and normalized axial stress (■) and the
computing time (□) in MPFEM (Procopio and Zavaliangos 2005).

2.3 3D MPFEM modelling


2.3.1 Modelling of single particle compression

Jerier et.al (2011) created a 3D finite element mesh of sphere with the aim of simulating
the particle in powder compaction. It was comprised of geometric simplifications for
MPFEM modelling, and the elements inside the sphere were coarser than surface area to
have a reasonable calculation time.

An additional type of 3D mesh is quadrangular elements. In 2007, Chen et al. (Chen et


al. 2007) studied a sphere compressed by two rigid planes to figure out the effect of
friction between the particle and the boundaries. The mesh of undeformed sphere after
14
compression and the outcomes of displacement-force curves were shown in Fig. 2.9.
The friction only affected the forces on punch after the displacement was large. In this
result, the final force without friction was a little smaller than the final force with
friction when the displacement was largest. Compared with experimental data, the
simulations showed a good agreement before the displacement of 6mm; only after
displacement was over 6mm, the force in experiment was a little smaller.

8000
Experimental data
3D, µ=0
6000 3D, µ=0.25
Force (N)

4000

2000

0
0 2 4 6 8 10
Displacement (mm)

Fig. 2.9 Comparison of the numerical results with two friction coefficients and
experimental results. (Chen et al. 2007).

Compression of single and two particles were modelled to test mesh sensitivity. Fig.
2.10a showed that the displacement-force curves were similar with different mesh
elements ranging from 425 to 2783, indicating that the results were mesh independent.
As the material modelled was a hyper-elastic polymer, the simulation results were
comparable with the results reported by Lin and Chen (2006) but different from
Hertzian force which describes the elastic deformation. Similar results were also
observed for the compression of two particles, as shown in Fig. 2.10b.

15
30

Force (mN)
20
425 elem.
1280 elem.
2783 elem.
10 Lin and Chen

0
0 0.1 0.2 0.3 0.4
Total displacement (mm)

(a)

30
Force (mN)

20

425 elem.
10 1280 elem.
2783 elem.

0
0 0.2 0.4 0.6 0.8
Total displacement (mm)

(b)

Fig. 2.10 Mesh sensitivity tests of (a) one particle and (b) two particles (Frenning 2008).

Harthong et al. (2009) reported the force-displacement behaviours of hydrostatic


compaction, die compaction and simple compression of a single particle. Due to the fact
that one sphere particle is symmetrical, only 1/8th of the particle was simulated. Fig.
2.11 indicates that the force on the boundary increases as the displacement becomes
larger and tends to reach infinity because the free space is smaller. A relation of force
and relative density can be summarized as follows: F→∞ when ρ→1. The figure

16
showed that the forces for different loading types depended not only on the
displacement of punch but also how much free space available. For the hydrostatic
compaction, as the displacement increased, the three contact areas increased as well. In
addition, when the three contact areas affected each other, overlaps took place. Since
then, the relationship between boundary displacement and force on boundary changed.
However, in a single particle die compaction and simple compression, the problem did
not occur. The bulk of the particle could extend left and right so that the force was
lower with the same displacement.

5
4.5 Die compaction

4
Hydrostatic
3.5
compaction
3
F/K2R2

2.5
2
1.5
1 Simple compaction
0.5
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
h/R

Fig. 2.11 Comparison of the force-displacement behaviour of hydrostatic compaction,


die compaction and simple compression (Harthong et al. 2009).

2.3.2 Modelling of multiple particle compression

Chen et al. (2007) studied a nine-particle cubic packing. In their study, the particles
were placed in a cylindrical die with fixed lower punch and upper punch compressed
downward. Fig. 2.11 showed the force-displacement curves with different friction
coefficients. The insets in the figure showed the initial and final states of the compact.
The closest results to the experimental data was the case of the friction coefficient: μSD
= 0.25 and μSS = 0.1. Where μSD means the friction coefficient of sphere and die, μSS
means the coefficient of sphere and sphere. There was a slight increase after a 10mm
displacement from the curve. This was because the upper four spheres firstly contacted
17
the lower four spheres. The friction in the packing system had obvious effects on
powder compaction, and the increase of friction increased force on the punch.

80000
Experimental data
µSD = 0, µSS = 0.05
60000
µSD = 0.1. µSS = 0.05
Force (mN)

µSD = 0.25, µSS = 0.1


40000

20000

0
0 5 10 15 20 25
Displacement (mm)

(c)

Fig. 2.12 Comparison of simulated and measured force-displacement. The insets


showed the initial and final states of the compact.

Jerier et.al. (2011) investigated the force law at high density. They simulated isostatic
and die compaction with both DEM and MPFEM and picked the lead and copper
parameters to make comparisons with the experimental results from the literature
(Martin et al. 2003). There was no friction between the spheres. They used the same
initial packing of 32 particles of radius is 0.15mm that were in a cube of 1 mm side. The
number of particles was controlled to have a reasonable calculation time for MPFEM.
For isostatic case, six box walls were moved inward at the same velocity. The data of
relative density and pressures on each wall were collected. From the figure we can see,
MPFEM showed a good capability of present deformation of particles in comparison to
DEM. The shape as well as the interaction of particles was monitored.

In the case of isostatic compaction, each wall moved inward at the same velocity. The
pressure on each wall was similar with relative density increasing. Fig. 2.13 showed the
stress–density curves of MPFEM and the comparison with DEM results and Storåkers’s
model. The high-density model is the DEM model applied a high-density contact law
from (Jerier et al. 2011) to describe interparticle contact at high relative density and
18
Storåkers’s model is a plastic model with a force law. The results of high density model
in DEM using new contact law were consistent with MPFEM. However, for Storakers’
model, the results were incorrect when the relative density was higher than 0.85. Fig
2.13a showed the stress-density curve by MFPEM, high density DEM model. The good
agreement between MPFEM and DEM result with high density contact law showed that
the contact law was appropriate for describing particle contacts in DEM when the
material was lead. Fig. 2.13b showed stress-density curve of MPFEM and the
experiment results from James (James, 1977). The results indicated that the MPFEM
results agreed with the experiment. On the other hand, Storakers’ model did not match
with the experiment result when the density was higher than 0.85.

40
35 High density model
30 MPFEM
Storakers' model
Stress (MPa)

25
20
15
10
5
0
0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
Relative density

(a)

19
500
High density model
400 Experimental result
Storakers' model
Stress (MPa) 300

200

100

0
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
Relative density

(b)

Fig. 2.13 Comparison between the results of DEM with new contact law, the MPFEM
results and the Storakers's model result (a) lead parameters; (b) MPFEM result with
copper parameters and experiments (Jerier et al. 2011).

2.4 Summary

The fundamentals of MPFEM and the applications of the model to powder compaction
have been reviewed. The information of particle stress and deformation in different
structures and different punch speed could be well known. The effect of particle size
and particle number were analysed using stress distribution and deformation distribution.

As the computer technology improved, 3D models of MPFEM were applied. While


number of particles were limited by simulation time, a basic understanding of particle
levels of particle compression was available. The force law that was used in high
density model was investigated using MPFEM. This force law can describe the particle-
particle force when the packing was in high density.

In conclusion, the analysis presented above demonstrated that DEM models is a suitable
way to simulate compaction behaviour at high relative density.

The objective of the current research is to produce a fundamental and feasible


understanding of powder compaction. MPFEM is an efficient approach for exploring
microcosmic mechanism in the process of powder densification in cold die compaction.

20
It can still be used to study the influence of particle microstructure in the macro-
structure of powder compacts. There is no systematic result on micro and macro
mechanism. The impact of particle size, three-dimensional, material properties and
inter-particle friction should be presented and the data of simulation will be compared
with the experimental data in order to achieve a better understanding of the compressing
process. While particle hardness is an important property in powder compaction and
may potentially affect particle arrangement and hardness of final packs, there is a gap in
the literature on particle hardness using MPFEM.

The specific objectives of this research are as follows:

• Develop an MPFEM model of high density powder compaction

• Validate the MPFEM model by comparing with previous work.

• Investigate the force and structural properties of compacts in compaction.

• Study the effects of particle hardness.

21
Chapter 3 MPFEM model development and validation

3.1 Introduction
Powder metallurgy (PM) is the process of making materials or components from metal
powders (German 1984). It can greatly reduce the loss of materials and be applied to a
variety of metals. The advantages of PM include easy operation, environment friendly
and near nett shape forming. Therefore, PM has been widely used in many industries
such as manufacturing, the automobile industry and the military industry.

Powder compaction is an important process in PM. During this process, loose powders
are compressed into to a coherent, dense state using tools such as punches, dies and
compression rolls (German 2005). During compaction, ductile particles experience
large plastic deformation and final relative density of the compact can be over 0.95. The
mechanical strength also increases. As the compact structure and strength affect the
following sintering process and the property of final product, it is crucial to understand
the dynamic behaviour of powders during compaction.

Considering the rapid development of computing technology, various numerical models


have been developed at different length and time scales in order to investigate the
compaction of powders. Models based on FEM have been designed (Sokolnikoff and
Specht 1956). In the models, a compact is treated as a porous visco-plastic material
governed by constitutive laws. At the particle scale, DEM is probably the most
commonly used approach (Cundall and Strack 1979). DEM is a fully dynamic method
to explicitly consider the inter-particle forces particle sizes, and material properties
(Aydin et al. 1997). Nevertheless, both approaches do not consider the large
deformation of powders at large densities. When the density is larger than 0.85, the
inter-particle contacts start to overlap, causing different interactions among the particles.
At present, there is no law describing the 3-body interaction. Therefore, a new
numerical approach is required to simulate high density compaction of powders.

Recently, the multi-particle FEM (MPFEM) method has been developed to consider the
change of particles during compaction (Procopio and Zavaliangos 2005). In MPFEM,
Particles are individually described with finite element mesh so that detailed
information on the particle mechanisms can be provided. MPFEM has been used to

22
simulate 2D (Zhang, 2015) and 3D compactions (Frenning 2008). The results on single
elastic sphere showed a good agreement with analytical expression using the Hertz
equation (Procopio et al. 2003). The MPFEM simulations also showed variation of
internal structure and stress state under different stain states (Procopio and Zavaliangos
2005). Chen et al. (Chen et al. 2007) compared MPFEM simulations of single particle
and a BCC packing of nine particles with experimental results. MPFEM could
accurately reproduce the behaviour of lead spheres. Moreover, the effects of the friction
between spheres and between sphere and boundary were studied. Harthong et al.
(Harthong et al. 2009) used MPFEM to obtain an analytical expression for the contact
force between particles and applied the equation in DEM to simulate high density
compaction. Zhang (Zhang, 2015) studied the compaction of compacts combined with
soft and hard particles using MPFEM. The effect of friction and volume of hard
particles is forced. They compared the compaction of discrete random packing and
several packings (tetragonal and hexagonal close packed structures).

In this chapter, we will develop a 3D MPFEM model to simulate the compaction of


powders. The model will initially be validated by comparing the results on single
particle and a BCC ordered packing with previous work (Chen et al. 2007). Thereafter
the model will be applied to the compaction of a random packing in a die. The
macroscopic response of pressure-density and internal structure are evaluated.

3.2 Model development


3.2.1 Constitutive model in MPFEM

The MPFEM model was developed using a commercial ABAQUS software. In the
model, the deformation of a particle can be divided into an elastic part and a plastic part,
given by:

ε = εel + εpl (3.1)

where ε is the total strain, εel is the elastic strain and εpl is plastic strain (Aydin et al.
1997).

The elastic deformation is characterised by Hooke’s law, given by (Aydin et al. 1997):

σel = Eεel (3.2)

where σel is elastic stress, E is the modulus of elasticity and εel is elastic strain.

23
In this work, the plastic model used by Chen et al. (Chen et al. 2007) was used to
describe the plastic deformation of particles, given by:

σ = K(εpl )n + σY (3.3)

where σ is the total stress, σY is yield pressure. K and n are parameters whose values
can be obtained from experiments (Chen et al. 2007).

3.2.2 Simulation conditions


Three cases were simulated in this work, compression of single particles, compaction of
a BCC ordered packing in a cylinder and compaction of a random packing in a cylinder,
as shown in Fig. 3.1. The first 2 cases were used to validate the MPFEM model while
the 3rd case was used to investigate the change of packing and stress structures during
compaction.

(a)

(b) (c)

Fig. 3.1 Compactions of powders: (a), single particle compression; (b), BCC ordered
packing; and (c), random packing

24
In the first case (Fig. 3.1a), a sphere of 18.48 mm was compressed using two rigid
planes moving at a constant speed of 10 mm/s till the displacement of a plane reached
10 mm. The meshes were generated by cutting the sphere into eight sections using the
three datum planes. The hexagon mesh was adopted in this study and the mesh size was
set at 0.8 mm. There were in 11,664 elements total and the element type was C3D8R.
Different sized meshes were also generated to test the mesh independence of the model.
The sphere-wall friction coefficient was 0.25.

In this study, the dynamic explicit method was used. In ABAQUS, there are two
available methods, namely: static analysis, which is used for stable problems including
linear and nonlinear response, and dynamic explicit method which uses relatively short
dynamic response times for large models. Because compaction is a dynamic process,
the dynamic explicit method is a suitable approach. The same method was also used for
other compactions.

The compactions of multiple particles in a cylinder (Fig. 3.1b and 3.1c) were modelled
by duplicating the single particle (e.g. size, material properties) and moving the
duplicated particles to various positions. For a BCC structure compact shown in Fig.
3.1b, the initial positions of the 9 particles were determined mathematically. On the
other hand, the positions of 100 particles in a random packing (Fig. 3.1c) were obtained
from a DEM simulation with a packing density of 0.485. In the compaction, the upper
punch moved downwards at a speed of 10 mm/s until the packing density reached 0.9.
After reaching a displacement of 10 mm, the plane moves upward, and the assembly is
unloaded. The particle size was 18.48 mm in the BCC compaction and 10 mm in the
random packing compaction. The particle-particle and particle-wall friction coefficients
were 0.05 and 0, respectively.

Table 3.1 lists some parameter values that were used in the simulation. The particles
had the material properties of lead. K and n are the fitting results from 10 relaxations
performed over a logarithmic strain ranging from 0.2 to 1.3 (Chen et al. 2007). Fig. 3.2
shows the comparison of the calculated stress-strain relations using Eq. 3.3 and the
experimental data. The results are in good agreement.

25
Table 3.1 Parameters used in present work.

Parameter Value
Young’s modulus, Y 10 GPa
Poisson’s ratio, 0.435
Punch speed, v 10 mm/s
yield strength, 𝝈𝒀 5 MPa
k 15.5
n 0.35

Fig. 3.2 Stress-strain relationship for lead spheres (Fitting using Eq. 4.2 with K= 15.5,
n=0.35) experimental results (Chen et al. 2007).

To automate the procedure, a Python script was written (see Appendix A), which was
then loaded from ABAQUS. The script was able to achieve the following functions: (i)
single particle creation and mesh generation; (ii) allocation of material properties to the
particle; (iii) reading particle positions from the position file (from mathematical
calculation or DEM simulation); and (iv) duplicating the original sphere with all the
information (size and material properties) and moving it to a position defined in the
position file. This step was repeatable till all the particles were generated and positioned;
and (v) defining cylinder position and material properties.

26
3.3 Results and discussion
3.3.1 Model validation
In this section, the model will be validated by comparing it with a previous study (Chen
et al. 2007). The effects of mesh size, time step and punch speeds are tested in the
compression of single sphere.

(i) Compression of single sphere

Fig. 3.3a shows the force-displacement relations of the single particle compaction. The
overall displacement is the particle radius. The results from Chen et al. (2007) are also
plotted for comparison. The agreement is very good. The relationship is non-linear due
to the elastic-plastic property of the sphere. To normalise the results, the stress-strain
relation is plotted in Fig. 3.3b. The stress is the pressure on the contact boundary
calculated by dividing the force by the contact area. The growth of two force-
displacement curves firstly is getting slower, then after a straight increase, it is getting
faster.

(a)

27
(b)

Fig. 3.3 Comparison of the (a) force-displacement; and (b) stress-strain relations for
compaction of a particle (Chen et al. 2007).

In the simulations, the mesh size, time step and punch speed are varied to study their
effects on the force-displacement relation: mesh size 0.4-1.2mm (Fig. 3.4), time step 10-
5
to 10-6 s and punch speed 2.5 mm/s to 10 mm/s. Fig. 3.5 shows the force-displacement
relation subject to different conditions, indicating that these variables have no visible
effect on the compaction curve.

(a) (b) (c)

Fig. 3.4 Mesh design with element size of (a) 1.2mm; (b) 0.8mm; (c) 0.4mm.

28
8000

7000 1.2 mm
0.8 mm
6000 0.4 mm
5000
Force (N)

4000

3000

2000

1000

0
0 2 4 6 8 10 12
Displacement (mm)

(a)

8000

7000 1e-05 s
5e-06 s
6000
1e-06 s
5000
Force (N)

4000

3000

2000

1000

0
0 2 4 6 8 10 12
Displacement (mm)

(b)

29
(c)

Fig. 3.5 Force-displacement curves under different conditions: (a) mesh sizes; (b) time
steps; and (c) punch speeds.

Fig. 3.6 shows the comparison of the final shape of the sphere and Von Mises stress
distribution. Von Mises stress σv is also referred to as equivalent tensile stress. It is
used to predict yielding of materials under multiaxial loading conditions, given by:

(σ1 −σ2 )2 +(σ2 −σ3 )2 +(σ3 −σ1 )2


σv = √ (3.4)
3

where σ1 , σ2 , σ3 are the uniaxial stress in three directions. The minimum stress occurs
as a cycle on the pressing surface (Hill, 1998). The figure shows that the current
simulation agrees well with previous study conducted by Chen et al. (2007).

30
(a) (b)

Fig. 3.6 Von Mises distribution (a) sphere compressed in simulation; (b) sphere
compressed in (Chen et al. 2007).

Fig. 3.7 is the Mises stress and pressure distribution of the centre section of the sphere.
In Fig. 3.7a, the Mises stress mainly occurs near the contact area and the maximum
stress is under the surface of contact. A similar trend is seen in Fig. 3.7a’ where the
maximum pressure exists at the contact surface. As the displacement is increased, an “H”
shaped high Mises stress area (Fig. 3.7b) demonstrates the maximum stress of the
particle occurs in the centre of the particle. In Fig 3.7b’, the largest pressure forms a
circle on the surfaces of upper and lower contact area. When the displacement is the
largest, Fig.3.7c shows that the ‘H’ shaped area extends to be larger. In Fig. 3.7c’, the
maximum pressure area shifts from the side cycle of contact surface to the centre of the
contact area. The main deformation takes place firstly at the contact surface when the
displacement is small. As the displacement is growing, the main deformation area
transfers to the centre of the particle. As the upper boundary is compressed, the
maximum pressure transfers to the centre of the contact surface and the area below the
centre of the contact surface.

31
(a) (a’)

(b) (b’)

(c) (c’)

Fig. 3.7 Mises stress and pressure of one particle compressing at different
displacements Mises stress: (a) 1mm; (b) 6mm; (c) 10mm. Pressure: (a’) 1mm; (b’)
6mm; (c’) 10mm.

Fig. 3.8 shows the average stresses along the horizontal (XY) and vertical (Z) lines
passing through the centre of the particle. LH and LV are the horizontal and vertical
distance from the centre of the particle, respectively. Fig. 3.8a shows the stress
variations along the horizontal direction. Note the negative signs in the figure means the
tensile stress. As shown in Fig. 3.8a, σxy maintains stable near the centre of the particle
and then decreases to a minimum (5 Pa) at the 2/3 of the particle radius and then
increases to near 15 Pa at the surface of the particle. Meanwhile, σz decreases from the
centre of the particle at about 37 Pa to almost zero at the surface of particle. Fig.3.8b
shows the stresses along the vertical line. The vertical stress σz increases slightly from
11.5 Pa to 16.8 Pa in the middle of the curve. σ xy remain stable (34.1 Pa) at first and
decrease to 32 Pa after a slight increase.

32
5
σxy
0
σz
-5

-10
Stress (MPa)

-15

-20

-25

-30

-35
0 2 4 6 8 10 12 14
LH

(a)

-5
σxy
-10
σz
Stress (MPa)

-15

-20

-25

-30

-35

-40
0 1 2 3 4 5
LV (mm)

(b)

Fig. 3.8 Average stress on the lines passing through the particle centre: (a) horizontal
line, and (b) vertical line.

33
(ii) Compaction of the BCC ordered packing

Fig. 3.9a shows the simulated force-displacement result for the BCC ordered packing.
The result from with previous work (Chen et al. 2007) is also plotted for comparison.
The current result shows a good agreement with the previous results. The compaction
curve suddenly increases near the 10mm displacement. This is because the contact of
four upper spheres and four lower spheres occurs. Fig. 3.9 (b) show the relationship
between pressure applied on the upper punch and relative density. The total trend is
similar with the force-displacement curve. The maximum pressure reaches 31.93 MPa
when the relative density is nearly 0.98.

(a)

34
(b)

Fig. 3.9 (a) Comparison of the pressing force of simulation and Chen et al. result (Chen
et al. 2007); (b) Pressure-strain results of BCC ordered simulation.

Figs. 3.10 shows the final state of the BCC compact. The result from the previous study
is also presented for comparison. The shapes of two compacts are nearly the same. The
Mises stress distributions are also similar. The average stress on the contact surface of
the particles with the punches is bigger than the contact surface of particle-particle and
particle-wall. The shapes of all eight corner particles are practically the same. Figs.
3.10a’ and b’ compare the shapes of the centre particle of the BCC packing from the
current work with the previous study. Both shape and stress distributions are similar.

(a) (b)

35
(a’) (b’)

Fig. 3.10 The final state of compaction (a) the current study; (b) study of Chen. Centre
sphere of (a’) 9-particle packing; (b’) simulation results from Chen. Figures b and b’ are
reproduced from Ref. (Chen et al. 2007).

3.3.2 Compaction of a random packing


Fig. 3.11 shows the variation of the pressure on the upper plate with packing density for
the duration of the compaction of the random packing. The initial density is 0.48, and
the density increases with the increases of pressure and drops with the decreasing of
pressure in the unloading step. Boundary pressure has a predictable increase as the
relative density increases. Once the relative density reaches 0.9, the unloading occurs,
and the pressure suddenly drops to zero with a little decrease of relative density.

18
16 C

14
loading
Pressure (MPa)

12
10
B
8
6
4 A unloading
D
2
0
0.45 0.55 0.65 0.75 0.85 0.95
Relative density

Fig. 3.11 Pressure as a function of density.

36
Four stages as marked in Fig. 3.11 have been selected for further analysis. Stages A, B
and C represent the beginning, middle and end of loading stage, stage D represent the
end of unloading.

Fig. 3.12 shows the stress distributions of the compact at four stages. The Mises stress
appears at the top of packing and numerous particles close to the top particle at first
(stage A, relative density is 0.55). After the particle rearrangement at point B, the stress
in the contact area of particle and die wall increases faster. After the relative density
reaches 0.9 (point C), the unloading starts. So the stress of the particles decreases
noticeably, as shown in Fig. 3.12D.

(a) (b)

(c) (d)

Fig. 3.12 Stress distributions at different stages of the compaction of the random
packing: (a) density = 0.55; (b) density = 0.7 (c) density = 0.9 (d) density ≈ 0.9.
37
The forces between the particles increases when the overlap between two particles
increases. Overlap is the extent to which the two particle centres get closer after
compression. It can be described as follows:

do = (R1 + R 2 ) − d (3.5)

where do is overlap, R1 and R 2 are radius of particles, d is distance between two


particles centre.

The distributions of the overlaps at four stages as illustrated in Fig. 3.13 (φ is the
overlap normalised by the particle diameter do/d). At stage A, the overlap ratio is mostly
all smaller than 0.1, and the probability increases till the ratio is 0.08 and then drops.
That means at this stage, all the contact overlaps are small. At stage B with relative
density of 0.7, the total contact number is 320 (more than stage A), The probability
increases slightly to peck when the ratio is 0.13 and then it decreases. The contact
number of particles reaches 352 at stage C. Hence, the largest overlap that the ratio is
over 0.4 exists at this stage. After unloading, with a little decrease of relative density,
the probability of small overlap (ratio smaller than 0.1) slightly increases in comparison
to stage C. The contact number of stage D is 350. The maximum overlap is smaller as is
expected. As the relative density increases, the number of contacts and overlaps also
increase before the unloading step.

1E+02

A
Probability density function P(φ)

1E+01
B

1E+00

1E-01

1E-02
0 0.1 0.2 0.3 0.4 0.5
overlap ratio,φ

Fig. 3.13 Distributions of particle overlaps at different stages.


38
Fig. 3.14 shows the distribution of force angles in compaction, which is the angle
between the force direction and the vertical direction (a vertical force has an angle of
zero). The figure shows that the force angles are predominantly in two areas, less than 5
degrees and near 25 degrees at stage A. This can be attributed to the compression
merely being the start with there being some room for the particles to rearrange
themselves. After the rearrangement, the force angles of last three stages all begin with
a sharp drop and then maintain a slight and fluctuant decrease to 0 which means force
angles nearly vertical is the most and the contact nearly horizontal is much less.

0.045
0.04 A
0.035 B
C
Probability density

0.03 D
0.025
0.02
0.015
0.01
0.005
0
0 20 40 60 80 100
Contact force angle, θF (deg)

Fig. 3.14 Force angle distribution of four stages in simulation of random packing.

Fig. 3.15 shows the number of particles on how many contacts with other particles. At
the beginning of compaction, most of the particles have 3-6 contacts with other particles
and no particle has contact number more than 8. At stage B, the particles have more
contacts and there is even a particle with 11 contacts. Up until the end of the loading,
particles which have 9 contacts visibly increase in comparison to stage B. Following the
unloading, the contact number become less and the particles with large contact numbers
decrease. In general, the result is reasonable to show the contacts increase as the relative
density increases.

39
Fig. 3.15 Curve of coordination number and number of particles.

The magnitude of force between particles is related to the overlap shown in Fig. 3.16.
The results were obtained from one particle compressed by two rigid planes. According
to the figure, there is a little fluctuation when the overlap is small, that is because elastic
deformation can be ignored. When the overlap is larger, the elastic force is too small in
comparison to the plastic force. Fitting line is added to the figure. The equation of
fitting line is:

Fc = 11222φ3 - 6495.9φ2 + 3681.2φ (3.6)

where Fc is contact force between two particles, φ = d/dp, d is displacement of the plane,
dp is the diameter of particles.

40
Fig. 3.16 Relationship of overlap and magnitude of force.

Fig. 3.17 shows the overlaps between the particles and the relationship of overlap and
magnitude of force. From the curves of four stages, it has a similar trend when
compared to overlap distribution. At stage A, the forces are near 50N and reach the
peak when the force is 51N. The largest force can only reach 67N. When comes to stage
B, the particle rearrangement is finished. The probability increases slightly to peck (near
8) when the force is 97N and then it decreases. Largest force increases from 67N in
Stage A to 195N. Forces probability densities in stage C and D are similar only the
probability is smaller in stage D when the force is near 290N. Compared to stage B,
force number is increased, and the increase mainly focus on number of forces larger
than 150N. The largest magnitude of force increases to 324N in stage C, then decreases
to 319N in stage D.

41
Fig. 3.17 Distributions of the contact forces at different stages.

3.4 Conclusion
In this chapter, a MPFEM model was developed. An elastic-plastic constitutive model
was adopted. The model was firstly used to model the compression of single particle
and a BCC order packing. The results were comparable with previous study so the
model was validated. Then the compaction of a random packing was simulated. In order
to figure out the interparticle forces, overlaps and force angle distributions were
calculated and analysed. The main findings can be summarised as follows:

• The pressure applied on punch increases when the relative density is larger.
• As relative density increases, the overlaps between particles are getting larger and
greater in number and the particles are surrounded by more particles coordination
number increases.
• Force angle distribution shows that most of the forces that are nearly vertical.

42
Chapter 4 Effects of particle hardness on compaction

4.1 Introduction
In the compaction process, hardness is one of main properties which affect the force and
arrangement of particles (Zhu et al. 2008). For example, the study of compaction
behaviour of four ceramic powders with different hardness showed the softer particles
yielded a larger volume compaction when subjected to pressure (Cooper and Eaton
1962). Cao et al. (2010) measured the particle hardness of different pharmaceutical
materials and they found that the materials with very low or high hardness were more
likely to have poor compaction behaviour.

In the iron and steel making process, ore and coke particles become soft at high
temperature. The softening of particles has significant effects on porosity and
permeability of packed beds, which affects the efficiency of the blast furnace. Therefore,
particle softening has been studied extensively to understand the behaviour of materials
during softening. Iwanaga (Iwanaga, 1989) applied load stress on burden materials at
high temperature to improve the understanding of softening behaviour of sinter. The
behaviour was evaluated by testing the gas pressure and the viscosity changes.
Nevertheless, since the high temperature experiments were difficult to conduct and the
results were affected by many factors, the effect of each factor on material soften was
not easy to be quantified.

Chew (Chew, 1999) developed a low temperature experiment to simulate the softening
and melting behaviour of sinter in the blast furnace. Paraffin wax balls which is a
material with a low melting point were used. A further study was undertaken by
Maldonado (Maldonado, 2003) who used wax balls as iron balls and glass beads to have
a better understanding of layered structure of cohesive zone. However, the information
of the impact on porosity due to individual factors was lacking. In light of this, Zheng
(Zheng, 2007) investigated the effects of temperature and load pressure on softening
behaviour of a packed bed using wax balls. In this study, the porosity changes of bed
due to particles deformation and permeability were examined subject to specified
temperature and load. It was found that the increase in temperature and load pressure
increased the deformation rate. The effect of temperature was the softening of wax balls
and the effects of load stress was through compaction due to ball deformation.
43
On the other hand, mixtures of powders with different mechanical features are
commonly used in powder compaction. These mixtures contain particles of different
hardness. To better understand the micromechanics and deformation in mixtures and
optimize the composition, it is important to study the densification behaviour of
compaction mix with soft and hard particles. Lange et al. (1991) studied powder
compaction with a mix of hard and soft particles. The soft particles played a greater role
in deformation and filling the gap of packing. Hardness affected both pressure between
particles and rearrangement of packing densification. Bouvard (2000) studied the
compaction of mixed hard and soft particles. The results showed that when the yield
pressure of soft particle was too high for applied pressure, the hard particles took most
part in re-arrangement; and when soft particles were most responsible for deformation
and small hard particles impede densification.

Numerical modelling has also been developed to study the compaction of packing with
a mixture of soft and hard particles. Gethin et al. (2002) created a DEM model to
simulate powder compaction of a combination of soft and hard particles. Martin and
Bouvard (2003) simulated a packing of 30% hard particles and 70% soft particles using
the DEM. Their results indicated that hard particles supported more load and the soft
particles were deformed under the squeezing of hard particles. Furthermore, the friction
between hard-hard particles played a significant role in the re-arrangement of particles.
Zhang (Zhang 2009) created a 2D MPFEM model with a mix of hard and soft particles
that is characterised by finite element mesh. The results showed the hard particles
increased compaction force and affected the distribution of load.

The purpose of this work is to explore the effects of particle hardness at high density
compaction using the previously developed MPFEM. Two processes will be modelled:
particle softening and mixture of hard and soft particles. The macroscopic relationship
of pressure and density, contact forces and structures during compaction are analysed.

4.2 MPFEM model and simulation conditions


4.2.1 Governing equations
The total strain of a particle consists of the elastic strain and plastic strain, given by:

ε = εel + εpl (4.1)

44
where ε is the total strain, εel is the elastic strain and εpl is the plastic strain (Aydin et al.
1997). For elastic strain, Hooke’s law is used to calculate:

σel = Eεel (4.2)

where σel is elastic stress and E is the modulus of elasticity. The hardening expression
for plastic behaviour is given from Chen et al. ( 2007):

σ = K(εpl )n + σY (4.3)

where σ is the total stress, σY is yield stress. K and n are parameter factors.

4.2.2 Simulation conditions

• Effect of particle softening

The compactions of particles in a cylindrical die were simulated. The particle softening
effect was considered by varying the hardness of particles at different temperatures. The
hardness of particles was so small that its effect on particle packing process was also
significant, which cannot be ignored. So, the simulations started with a packing process
of 100 particles in which the consolidation of the particles was driven by the gravity
only. As mentioned in Chapter 3, the initial positions of the particles were generated
using the DEM. In the simulations, the coordinates of the particles were import to
Abaqus to generate the initial packing. After the packing process, the compaction
started by moving the upper punch downwards at a constant speed of 10 mm/s till a
prescribed density was achieved.

Table 4.1 lists the values of the parameters used in the simulations. In the simulations,
particle hardness at 3 temperatures (35 ℃, 37 ℃ and 39 ℃) were adopted. The hardness
E (Young’s modulus), constant K and yield stress were calibrated based on the
experimental results from Zheng (2007). Fig. 4.1 shows the stress-strain relation of
particles with different hardness.

Table 4.1 Parameters of in the particle softening simulations.

Parameter Base value at 35℃ (values at 37℃, 39℃)

Punch speed, v (mm/s) 10

Friction coefficient, μ 0.4


45
Time step (s) 5×10-6

Particle diameter, d (mm) 15

Poisson’s ratio, 𝜈 0.435

n 0.35

Young’s modulus (GPa) 40 (16.8, 6.8)

Stiffness K (N/mm) 0.005 (0.0021, 0.00085)

Yield stress, 𝜎𝑌 (KPa) 6 (2.52, 1.02)

0.014
35℃

0.012 37℃
39℃
0.01
Stress (MPa)

0.008

0.006

0.004

0.002

0
0 0.2 0.4 0.6 0.8 1
Logarithmic strain

Fig. 4.1 Strain-stress curve of different hardness.

• Effect of particle mixture

In the simulation of particle mixture, 50 hard particles and 50 soft particles were mixed
and compacted in a cylindrical die. Fig. 4.2 shows the initial configuration of particles.
During compaction, the upper punch surface moved downwards at a constant speed of
10 mm/s until a prescribed density was achieved.

46
Fig. 4.2 Initial configuration of the particles. Black particles are hard and white particles
are soft.

Table 4.2 shows the parameters in the hardness model. In the simulations, the hardness
of the hard particles was 10 times that of soft particles. The constant K and yield
pressure also varied in proportion. Fig. 4.3 compare the stress-strain relations for the
hard and soft particles.

Table 4.2 Parameters of hardness model.

Parameter Value

Punch speed, v (mm/s) 10

Friction coefficient, μ 0.05

Time step (s) 5×10-6

Particle diameter, d (mm) 10

Poisson’s ratio, 𝜈 0.435

n 0.35

Young’s modulus, (GPa) 10 (soft particles) and 100(hard particles)

Stiffness K (N/mm) 15.5 (soft), 155(hard)

Yield stress, 𝜎𝑌 (MPa) 5 (soft). 50 (hard)

47
100
Hard particles

Soft particles

10
Stress (MPa)

0.1
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Loarithmic strain

Fig. 4.3 Strain-stress relations of soft and hard particles.

4.3 Results and discussion


4.3.1 Effect of particle softening

• Particle packing

Particle packing is a dynamic process. To examine if a stable packing has been formed,
the total force on the die bottom surface was monitored, as shown in Fig. 4.4. The
particles fall down under gravity and when impact with the bottom wall, their kinetic
energy is quickly dissipated. The total force on the bottom wall of 3 temperatures shows
similar trends of a damped oscillation which gradually stabilise after 0.3 s at around
1.55 N which is the total weight of the particles inside the die.

It is also observed that the curves of different temperatures are different in term of the
positions and number of the peaks. For the 35 ℃ case, the force climbs to a peak of 3 N
at around 0.05s. In the 37 ℃ and 39 ℃ cases, the time to have the 1st peak occurs at to
0.07 s and 0.08 s, respectively, indicating the time to reach the peak is larger for softer
particles. This is because the softer particles have larger deformation and require longer
time to become stable.

48
This is also reflected in Fig. 4.5 which shows the final states of the packings under
gravity. It is apparent that the packing of wax particles at 39℃ has the largest
deformation compared with other two packings. The softer particles at the bottom
deform significantly more at 39℃ than the other 2 cases.

3.5
35℃
3
37℃
Force on the die bottom (N)

2.5 39℃

1.5

0.5

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (s)

Fig. 4.4 Forces on the bottom wall for the packing of different temperatures (35 ℃, 37 ℃
and 39 ℃).

(a) (b) (c)

Fig. 4.5 Final states at different temperatures after packing under gravity: (a) 35℃; (b)
37℃; (c) 39℃. The colour represents the magnitudes of the normal forces.

49
Fig. 4.6 shows the distributions of the overlap between the particles. At the lower
temperature with harder particles, the distribution is narrower. For example, the largest
overlap in the 35℃ case is 0.12d (d is particle diameter) with a peak at 0.08d. In the 37 ℃
case, the peak position is at 0.1d and the maximum overlap is 0.175d. In the 39 ℃ case,
the distribution is flatter with a peak at 0.11d and the maximum overlap of 0.26d. It is
clear that the hardness of the particle affects the overlaps between particles in forming a
packing under gravity.

25
35℃
Probability density function P(φ)

20 37℃

39℃
15

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3
φ (overlap ratio)

Fig. 4.6 Distributions of the inter-particle overlap in three packings.

Force direction is another important property to characterise inter-particle forces Fig.


4.7 shows the distribution of the angles between the forces and gravity direction. The
highest probability density of three curves are all at the area of force angle from 0
degree to 5 degrees. Then, the probability densities fluctuate until the force angle is 50
degrees. Then the densities decrease after the angle is larger than 50 degrees. Therefore,
the forces which are nearly vertical have the largest number and the forces which are
nearly horizontal are rare.

Coordination numbers, which is the number of particles that contact with an object, is a
fundamental feature of packings and impacts the pressure transport. Fig. 4.8 shows the
distribution of coordination numbers. In the 35 ℃ case, the peak position is at 5 (29%).
50
With increasing temperature and decreasing hardness of the particles, the curve moves
to the right with larger coordination number. The peak positions for the 35 ℃ and 37 ℃
cases are 6 and 7, respectively. This indicates that softer particles have more contacts
with neighbouring particles because of larger deformation.

0.04
35℃
0.035
Probability density function P(θ)

37℃
0.03
39℃
0.025

0.02

0.015

0.01

0.005

0
0 20 40 60 80 100
θ (deg)

Fig. 4.7 Probability density and force angle in three packings before compression.

35%
35℃
30%
37℃
25% 39℃
Percentage

20%

15%

10%

5%

0%
0 5 10 15
Coordination No., Cn

Fig. 4.8 Distribution in coordination number for three packing before compression.

51
• Compaction of particles

After the initial packings are formed, compactions are conducted and the results can be
discussed. Fig. 4.9 shows the variations of porosity with compaction pressure at
different temperature. For comparison, the experimental results (Zheng 2007) are also
plotted. The porosities of the compacts decrease exponentially when the load pressure
on punch increases. Due to the fact that the hardness of particles at different
temperatures are different, the initial porosities that are 0.45, 0.41 and 0.35 for 35 ℃ 37 ℃
and 39 ℃ models respectively are different when subjected to the effects of gravity. The
results of MPFEM at 35 ℃ and 37 ℃ are in consistent agreement with the experimental
results. However, the porosity at same pressure in the simulations is slightly little higher
than the experiment results. This is because the structures of the two packings are
different and the initial porosities are different. At the low pressure in the 39 ℃ case,
the porosity of numerical packing shows a higher result than the experimental one.
Nonetheless, the simulation results are qualitatively agreeing with the experimental
measurement.

0.5
0.45 Exp-35℃
MPFEM-35℃
0.4
Exp-37℃
0.35 MPFEM-37℃
0.3 Exp-39℃
Proosity

MPFEM-39℃
0.25
0.2
0.15
0.1
0.05
0
0 1 2 3 4 5 6
Pressure (KPa)

Fig. 4.9 Fitting results of final porosity and pressure on punch for experiments (Zheng
2007) and simulations in different temperature. The line indicates shows the compact
structure at the pressure of 1.2KPa

52
To examine the effect of temperature on compact structure, three compacts obtained
with the pressure of 1.2 KPa are selected for analysis. The porosities of three compacts
are 0.31, 0.21 and 0.02, respectively. Fig. 4.10 illustrates the packing structures and the
stress distributions of three compacts. It is evident that more deformation takes place
for soft particles. Also, the Mises stress is greater for harder particles. The total particle
arrangements of the three cases after compression are similar except a few particles at
the top of the compacts are different. This may be due to the difference of packing
velocity caused by gravity causing upper particles to have different locations.

(a)

(b)

53
(c)

Fig. 4.10 Compact structure at the pressure of 1.2KPa: (a) 35℃; (b) 37℃; (c) 39℃.

Fig. 4.11 shows the probability distributions of the overlap between the particles at
three temperatures. The particles are harder when the temperature is lower. At 35 ℃, the
peak is at the overlap of 0.12 and the maximum overlap can reach 0.22. While at 37 ℃,
the peak is at the overlap of 0.2 and the maximum overlap increases to 0.35. And the
largest overlap for the 39 ℃ case is at 0.45. The probability decreases when the overlap
is larger. The average overlap in the three cases increases with the temperature
decreasing. Subject to the same pressure, the packing with harder particles has smaller
overlaps since the relative density is smaller.

10
9 35℃
Probability density function P(φ)

8 37℃
7 39℃
6
5
4
3
2
1
0
0 0.1 0.2 0.3 0.4 0.5 0.6
φ

Fig. 4.11 Probability density function of overlaps in the compacts.

Fig. 4.12 shows the distribution of the angles between the forces and gravity direction in
three models. The force angles are primarily between 0 degree to 5 degrees which
means the largest number of angles are almost vertical. Then the probability density
decreases swiftly from near 0.032 to 0.014 when the angle is 5 degrees. The angle of
forces has a slightly increase during 5 to 50 degrees. The force angles in three models

54
shows the same trend. So, particle hardness has little effect on force direction during
powder compaction.

0.04
35℃
0.035
Probability density function P(θF)

37℃
0.03
39℃
0.025

0.02

0.015

0.01

0.005

0
0 20 40 60 80 100
Contact angle, θF (deg)

Fig. 4.12 Probability density and force angle in three packings after compression.

The average contact number in the 35 ℃ case is 5.72, 6.46 in 37 ℃ case and 6.92 in 39 ℃
case. Fig. 4.13 shows the number of particles and the number of contacts with other
particles in the three cases. For the 35 ℃ case, the number of particles increases as the
coordination number increases until it reaches a peak of 29, when the coordination
number is 6 then the percentage of particle decreases. For the 37 ℃ case, the coordinate
number varies from 2 to 12. For the 39 ℃ case, the peak shifts to the right with a peak
at 7. This indicates that softer particles have more contacts with neighbouring particles
because of larger deformation.

55
35%
35℃
30%
37℃
25%
39℃
Percentage

20%

15%

10%

5%

0%
0 5 10 15
Coordination No., Cn

Fig. 4.13 The distribution in coordination number for three packings after compression.

The inter-particle forces can be determined from the overlaps. Fig. 4.14 shows the
relationships between the force and overlap on compression of one particle using two
rigid planes. Based on the figure, the force-overlap relation is given by:

Fc = K x × φmx (4.4)

where Fc is the contact forces and φ is the overlap. K x and mx are the fitting parameters
and their values are listed in Tab.4.3.

Table 4.3 Fitting parameters of three models.

Case 𝐾𝑥 mx
35℃ 4.511 1.5225
37℃ 1.851 1.5056
39℃ 0.728 1.4862

56
1.8
35℃
1.6 37℃
39℃
1.4 Fitting 35℃
Fitting 37℃
1.2 Fitting 39℃
Force (N)

0.8

0.6

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6
do/dp

Fig. 4.14 Curves of force and overlap ratio in one particle compression with three
hardness and fitting curves.

Fig. 4.15 shows the normalized contact force distribution in three cases calculated by
fitting equations and overlaps. All the forces are normalized by the averaged value. The
force distributions for the three cases are similar, all showing a trend of exponential
decay. In the 35℃ case, where the particles are hardest, P(Fc) exhibits a decay as the
forces get larger after a slight increase from 2.1 to 3.7 when the fc is less than 0.1. In the
37℃ case, the distribution increases from 2.2 to 5.3 before Fc is 1.4, and then decrease
to 0.26 when Fc is 3.7. In the 39℃ case, the distribution has a little fluctuation till Fc is
1.4 and then falls to 0.27.

57
10
35℃
Probability density function, P(Fc)
37℃

39℃

0.1
0 1 2 3 4 5
Fc

Fig. 4.15 Distribution of normalized contact force between particles in models with
different hardness under same pressure.

4.3.2 Effect of particle mixture


In this work, 100 particles of different hardness are mixed and compacted. While the 50
hard particles have the same properties as those in Chapter 3, the other 50 soft particles
are 10 time softer. They are compacted to show the effect of soft particles on
compaction behaviour.

Fig. 4.16 compares the pressure-density relation of two compacts. Both compacts have
the same initial relative density as they start from the same structure. For the hard
particle compact, the pressure increases in a much faster rate than that of the particle
mixture. The pressure required to compact the hard particle compact to 0.9 is about 5
times the pressure required to compact the mixture compact to the same density. This
indicates the macroscopic responses of particle mixture can be predicted from the
compaction of particles of the same properties. However, the unloading curves of two
cases are almost the same, indicating that the elastic recovery is mainly dominated by
the hard particles.

58
18

16 Hard particle compact


14 Mixture compact

12
Pressure (MPa)

loading
10

8
unloading
6 C
4 A B

2
D
0
0.45 0.55 0.65 0.75 0.85 0.95
Relative density

Fig. 4.16 Curve of relative density and upper punch pressure in hardness model
including loading and unloading.

Similar to the analysis in Chapter 3, four stages with different densities (A, B, C, and D
in Fig. 4.15) are selected and analysed. The packing densities of the four stages are 0.55,
0.7, 0.9 and 0.89 respectively.

Fig. 4.17 shows the compact structure and Mises stress at different stages. At all stages,
the stress is primarily distributed among the hard particles. In the compaction, the
particle rearrangement is smaller compared to the hard particle compact due to the
larger deformation of the soft particles. When the compact reaches its maximum density
(Stage C), The stress is significantly larger than in the other stages. The soft particles
deform significantly and fill the space among the particles. At the end of unloading
(Stage D), the particles remain their shape largely.

Similar observations can be found in the pressure distributions at different stages, as


shown in Fig. 4.18. At stage A, the pressure on the particles is very small compared
with stages B and C since the forces between the particles are still small as the relative
density is low. The pressure at stage C is the largest and mostly exists at the particle-

59
particle and particle-wall contact areas. After unloading, the pressures between particles
drops abruptly.

(a) (b)

(c) (d)

Fig. 4.17 Compact structure and Mises stress distributions at different stages: (a) stage
A; (b) stage B; (c) stage C; (d) stage D.

(a) (b)

60
(c) (d)

Fig. 4.18 Pressure distribution of four stages in hardness model results. (a)stage A; (b)
stage B; (c) stage C; (d) stage D.

Fig. 4.19 shows the distributions of inter-particle overlaps at different stages. At stage A,
the overlaps are very narrow and smaller than other stages. However, they are larger
than in the hard particle compact due to large deformation of the soft particles even at
the early stage. At stage B, the distribution becomes wider and the peak moves to a
larger value, which is similar to the hard particle compact. The distributions of stage C
and D are similar showing much broader distributions.

40
A
35
Probability density function P(φ)

B
30 C
D
25

20

15

10

0
0 0.1 0.2 0.3 0.4 0.5
Overlap, φ (d)

Fig. 4.19 Distribution of overlap in hardness model.

Fig. 4.20 shows the distribution of the force orientation characterised by the angle
between the forces and the vertical direction. At stage A, a large amount of forces has

61
angles between 0 to 5 degrees. Then it decreases to about 0.013 till force angle is 47
degrees. After that, a peak appears at the area of 52 to 60 degrees. At stage B, the
maximum probability is similar to stage A however there is no peak at the same area.
For stages C and D, similar trends occur. The probability starts from 0.025 in area of 0
to 5 degrees and then decreases until 0 at 90 degrees. The main angle at all stages are
under 5 degrees similar with the hard particle compact. However, there is a visible
increase near 50 degrees at stage A. The reason is that less rearrangement leads to few
changes for angles of force between the particles.

0.035
A
Probability density function P(θF)

0.03 B
C
0.025
D
0.02

0.015

0.01

0.005

0
0 20 40 60 80 100
Angle of force, θF (deg)

Fig. 4.20 Force angle distribution of hardness model at four stages.

Fig. 4.21 shows the distributions of coordination number. At the beginning of the
compaction (stage A), the most probable coordination number is 5. With increasing
pressure and density, the peak positions move to 6 and 7 for stage B and C, respectively.
The distributions also become wider with more particles having more contacting
particles. The distribution of stage D shows a similar one as that at stage C, indicating
the compact structure remain unchanged after unloading.

62
35%
A
30% B
C
25% D
Percentage

20%

15%

10%

5%

0%
0 2 4 6 8 10 12 14
Coordination No., Cn

Fig. 4.21 Coordination number and particle numbers in hardness model

4.4 Conclusion

In the chapter, the previous MPFEM model has been adopted to study the effect of
particle hardness on high density compaction. Two cases were selected: softening
particles at higher particles and mixing of hard and soft particles.

In the study of particle softening, three cases at different temperatures were simulated.
The variations of particle hardness with temperature were calibrated from the
experimental data. To consider the effect of initial packing condition, both packing and
compaction were simulated. In the packing stage, the settling of the particles was due to
gravity. The forces on the bottom plate showed a damped oscillation before reaching a
stable value which was the total weight of the particles. The structure and forces in the
packings were analysed. The packings were then compacted under different pressures.
The simulation porosity-pressure relation showed a good agreement with the
experimental data. Under a same loading, the porosity of packings of the softer particles
was lower than the harder particles.

The mixture of soft and hard particles was simulated to investigate the deformation
behaviour of the mixture under pressure. The soft particles played a larger part in
deformation and the hard particles surrounded by the soft particles showed little
63
deformation. Nevertheless, the maximum stress occurred at the contact area of
boundary and hard particles, and the stress in the hard particles was larger than in the
soft particles. As the relative density increased, the average overlap between the
particles became larger and the average coordination number increased.

64
Chapter 5 Conclusions and future work

In this work, an MPFEM model using ABAQUS was developed to investigate the
micro-mechanics and structure of packings at high density. The model was validated
based on the compaction of single particle and ordered packing. The model was then
adopted to study the compaction of random packings. The effect of particle hardness
was also investigated. The key findings are as follows:

● In the compression of single particles, the current results were comparable with
previous study (Chen et al. 2007). The results also showed independence on the size
of mesh, punch speed and time steps. The analysis to the Mises stress and pressure
distribution showed an H-shaped area in the centre of sphere. The centre of the
particle-punch contact had the highest normal force.
● The simulation of the compaction a BCC ordered packing also agreed well with
previous study, further confirming the validity of the model. The force-displacement
curve was similar to the experiment results and the deformation and stress
distribution were nearly the same.
● In the compaction of a random packing, four stages were analysed representing the
beginning, middle and end of loading and unloading. The compact density increased
with pressure. Force angles were mainly near vertical and the probability density
decreased when the angles were larger at beginning. The average contact number of
particles increased as the relative density increases. Forces at high relative density
stage were noticeably larger than at lower relative density.
● Packing and compaction of wax particles at different temperatures were simulated.
The softening effect of particles at high temperature was considered by reducing
particle hardness at higher temperature. At the packing stage, the particles were
settled under gravity. The forces on the bottom surface were monitored to determine
when the packings were formed. The softer particles required longer time to settle
and formed a packing of smaller porosity. At the compaction stage, the simulated
results of porosity showed a good agreement with the experimental results. The
overlaps, force angles and coordination numbers, did not show significant changes
in the three cases, however the forces between particles were significantly larger in
the case with harder particles.

65
● To explore the deformation behaviour and inter-particle forces of mixture of soft
and hard particles, a model was created using MPFEM. Although, hard particles
surrounded by soft particles showed little deformation, the maximum stress was
observed at the contact area of the boundary and hard particles. Soft particles played
the greatest role deformation. The different rearrangements of the two packings led
to different force angle and coordination number distributions. The probability of
force angles near vertical of loading was visibly smaller than the beginning of
loading.

In this research MPFEM models were proved to be a suitable method for presenting
structure and stress in random packing. Nevertheless, the interactions were limited to
simple friction models and no cohesive behaviour was included. The effect of particle
hardness in powder compaction is explored however the inclusion of effect of particle
size which may affect arrangement of particles in packing is lacking. All of the particles
are treated as ideal spheres. While it is difficult to produce a mixture of different shape
particles like actual particles in powder, it is also necessary to include the effect of
particle shape. Moreover, only 100 particles were included in the random packing due
to computational limitations. In addition, it is possible to hold more particles in order to
study the impact of particle number when more reasonable time step and mesh size with
more developed computers comes up. The mixture model of this study contains only
two different hardness of particles. Therefore, more levels of hardness need to be
included.

Although, the information of interactions between particles could be extracted using


experimental methods like atomic force microscopy and nano-indentation, experiments’
measurements are restricted to small scale deformations and a few plastic deforming
contacts. Therefore, the inclusion of experimental measurements of particle-particle
interactions is necessary.

66
References
Abaqus Analysis User's Manual 6.14.
Aydin, I., B. Briscoe and N. Ozkan (1997). "Modeling of powder compaction: a
review." MRS Bulletin 22(12): 45-51.
Aydin, İ., B. J. Briscoe and K. Y. Şanlitürk (1994). "Density distributions during the
compaction of alumina powders: A comparison of a computational prediction with
experiment." Computational materials science 3(1): 55-68.
Bhatt, J., M. Carroll and J. Schatz (1975). "A spherical model calculation for volumetric
response of porous rocks." Journal of Applied Mechanics 42(2): 363-368.
Bouvard, D. (2000). "Densification behaviour of mixtures of hard and soft powders
under pressure." Powder technology 111(3): 231-239.
Cao, X., M. Morganti, B. C. Hancock and V. M. Masterson (2010). "Correlating particle
hardness with powder compaction performance." Journal of pharmaceutical sciences
99(10): 4307-4316.
Carroll, M. M. (1980). "Compaction of dry or fluid-filled porous materials." Journal of
the Engineering Mechanics Division 106(5): 969-990.
Chen, Y., D. Imbault and P. Dorémus (2007). Numerical simulation of cold compaction
of 3d granular packings. Materials science forum, Trans Tech Publ.
Cheng, Y., S. Guo and H. Lai (2000). "Dynamic simulation of random packing of
spherical particles." Powder Technology 107(1): 123-130.
Cooper, A. and L. Eaton (1962). "Compaction behavior of several ceramic powders."
Journal of the American Ceramic Society 45(3): 97-101.
Cundall, P. A. and O. D. Strack (1979). "A discrete numerical model for granular
assemblies." Geotechnique 29(1): 47-65.
Frenning, G. (2008). "An efficient finite/discrete element procedure for simulating
compression of 3D particle assemblies." Computer Methods in Applied Mechanics and
Engineering 197(49): 4266-4272.
German, R. M. (1984). "Powder metallurgy science." Metal Powder Industries
Federation, 105 College Rd. E, Princeton, N. J. 08540, U. S. A, 1984. 279.
German, R. M. (2005). Powder metallurgy and particulate materials processing: the
processes, materials, products, properties, and applications, Metal powder industries
federation Princeton, NJ.
Gethin, D. T., R. W. Lewis and R. S. Ransing (2002). "A discrete deformable element

67
approach for the compaction of powder systems." Modelling and simulation in
Materials Science and Engineering 11(1): 101.
Harthong, B., J.-F. Jérier, P. Dorémus, D. Imbault and F.-V. Donzé (2009). "Modeling of
high-density compaction of granular materials by the discrete element method."
International Journal of Solids and Structures 46(18): 3357-3364.
Harthong, B., J.-F. Jérier, V. Richefeu, B. Chareyre, P. Dorémus, D. Imbault and F.-V.
Donzé (2012). "Contact impingement in packings of elastic–plastic spheres, application
to powder compaction." International Journal of Mechanical Sciences 61(1): 32-43.
Jerier, J.-F., B. Hathong, V. Richefeu, B. Chareyre, D. Imbault, F.-V. Donze and P.
Doremus (2011). "Study of cold powder compaction by using the discrete element
method." Powder Technology 208(2): 537-541.
Johnson, K. L. and K. L. Johnson (1987). Contact mechanics, Cambridge university
press.
Kogut, L. and I. Etsion (2002). "Elastic-plastic contact analysis of a sphere and a rigid
flat." Journal of applied Mechanics 69(5): 657-662.
Kyung-Hun, L., L. Jung-Min and K. Byung-Min (2009). "Densification simulation of
compacted Al powders using multi-particle finite element method." Transactions of
Nonferrous Metals Society of China 19: s68-s75.
Lange, F., L. Atteraas, F. Zok and J. Porter (1991). "Deformation consolidation of metal
powders containing steel inclusions." Acta metallurgica et materialia 39(2): 209-219.
Lautenschlager, E. and J. Harcourt (1970). "Photoelastic observations in diametral
compression testing." Journal of dental research 49(1): 175-175.
Lin, Y. Y. and H. Y. Chen (2006). "Effect of large deformation and material nonlinearity
on the JKR (Johnson–Kendall–Roberts) test of soft elastic materials." Journal of
Polymer Science Part B: Polymer Physics 44(19): 2912-2922.
Mahoney, F. M. and M. Readey (1995). Applied mechanics modeling of granulated
ceramic powder compaction, Sandia National Labs., Albuquerque, NM (United States).
Martin, C. and D. Bouvard (2003). "Study of the cold compaction of composite
powders by the discrete element method." Acta Materialia 51(2): 373-386.
Martin, C., D. Bouvard and S. Shima (2003). "Study of particle rearrangement during
powder compaction by the discrete element method." Journal of the Mechanics and
Physics of Solids 51(4): 667-693.
Mesarovic, S. D. and N. A. Fleck (2000). "Frictionless indentation of dissimilar elastic–

68
plastic spheres." International Journal of Solids and Structures 37(46): 7071-7091.
Naylor, D. (1978). "Stress–strain laws for soil." Developments in Soil Mechanics–1.
Edited by CR Scott. Applied Science Publishers, Ltd., London.
Procopio, A., A. Zavaliangos and J. Cunningham (2003). "Analysis of the diametrical
compression test and the applicability to plastically deforming materials." Journal of
Materials Science 38(17): 3629-3639.
Procopio, A. T. (2006). On the compaction of granular media using a multi-particle
finite element model.
Procopio, A. T. and A. Zavaliangos (2005). "Simulation of multi-axial compaction of
granular media from loose to high relative densities." Journal of the Mechanics and
Physics of Solids 53(7): 1523-1551.
Ransing, R., D. Gethin and A. Khoei (2000). "Powder compaction modelling via the
discrete and finite element method." Materials & Design 21(4): 263-269.
Ransing, R., D. Gethin, A. Khoei, P. Mosbah and R. Lewis (2000). "Powder compaction
modelling via the discrete and finite element method." Materials & Design 21(4): 263-
269.
Shima, S. and M. Oyane (1976). "Plasticity theory for porous metals." International
Journal of Mechanical Sciences 18(6): 285-291.
Sinha, T. (2011). "Investigation of compaction, Young's modulus and tensile strength of
binary powder mixtures using the multi-particle finite element method (MPFEM),
Purdue University."
Sokolnikoff, I. S. and R. D. Specht (1956). Mathematical theory of elasticity, McGraw-
Hill New York.
Storåkers, B., S. Biwa and P.-L. Larsson (1997). "Similarity analysis of inelastic
contact." International Journal of Solids and Structures 34(24): 3061-3083.
Wu, C.-Y., C. Thornton and L.-Y. Li (2003). "Coefficients of restitution for elastoplastic
oblique impacts." Advanced Powder Technology 14(4): 435-448.
Zhang, J. (2009). "A study of compaction of composite particles by multi-particle finite
element method." Composites Science and Technology 69(13): 2048-2053.
Zhang, Y., X. An and Y. Zhang (2015). "Multi-particle FEM modeling on microscopic
behavior of 2D particle compaction." Applied Physics A 118(3): 1015-1021.
Zhu, H., Z. Zhou, R. Yang and A. Yu (2008). "Discrete particle simulation of particulate
systems: a review of major applications and findings." Chemical Engineering Science

69
63(23): 5728-5770.
Zienkiewicz, O. and Z. Mroz (1984). "Generalized plasticity formulation and
applications to geomechanics." Mechanics of engineering materials 44: 655-679.
Martin, C. and D. Bouvard (2003). "Study of the cold compaction of composite
powders by the discrete element method." Acta Materialia 51(2): 373-386.

70
Appendix A.Python script to generate a random packing in Abaqus
# -*- coding: mbcs -*-
from part import *
from material import *
from section import *
from assembly import *
from step import *
from interaction import *
from load import *
from mesh import *
from optimization import *
from job import *
from sketch import *
from visualization import *
from connectorBehavior import *

# define part, model, and material name"


part_org_name = "Sphere-"
part_name = part_org_name + "1"
model_name = "Model-1"
material_name = "Material-1"
np_total = 100
mesh_size = 0.8
particle_radius = 5
time_step = 5e-6
friction_coefficient = 0.05
Boundary_velocity = -10
Time_period = 4
# create sphere
mdb.models[model_name].ConstrainedSketch(name='__profile__', sheetSize=200.0)
mdb.models[model_name].sketches['__profile__'].ConstructionLine(point1=(0.0,
-100.0), point2=(0.0, 100.0))
mdb.models[model_name].sketches['__profile__'].FixedConstraint(entity=
mdb.models[model_name].sketches['__profile__'].geometry[2])
mdb.models[model_name].sketches['__profile__'].ArcByCenterEnds(center=(0.0, 0.0)
, direction=CLOCKWISE, point1=(0.0, particle_radius), point2=(0.0, - particle_radius))
mdb.models[model_name].sketches['__profile__'].CoincidentConstraint(
addUndoState=False, entity1=
71
mdb.models[model_name].sketches['__profile__'].vertices[2], entity2=
mdb.models[model_name].sketches['__profile__'].geometry[2])
mdb.models[model_name].sketches['__profile__'].CoincidentConstraint(
addUndoState=False, entity1=
mdb.models[model_name].sketches['__profile__'].vertices[0], entity2=
mdb.models[model_name].sketches['__profile__'].geometry[2])
mdb.models[model_name].sketches['__profile__'].CoincidentConstraint(
addUndoState=False, entity1=
mdb.models[model_name].sketches['__profile__'].vertices[1], entity2=
mdb.models[model_name].sketches['__profile__'].geometry[2])
mdb.models[model_name].sketches['__profile__'].Line(point1=(0.0, particle_radius), point2=(
0.0, - particle_radius))
mdb.models[model_name].sketches['__profile__'].VerticalConstraint(addUndoState=
False, entity=mdb.models[model_name].sketches['__profile__'].geometry[4])
mdb.models[model_name].sketches['__profile__'].PerpendicularConstraint(
addUndoState=False, entity1=
mdb.models[model_name].sketches['__profile__'].geometry[3], entity2=
mdb.models[model_name].sketches['__profile__'].geometry[4])
mdb.models[model_name].Part(dimensionality=THREE_D, name=part_name, type=
DEFORMABLE_BODY)
mdb.models[model_name].parts[part_name].BaseSolidRevolve(angle=360.0,
flipRevolveDirection=OFF, sketch=
mdb.models[model_name].sketches['__profile__'])
del mdb.models[model_name].sketches['__profile__']

# generate mesh
mdb.models[model_name].parts[part_name].DatumPlaneByPrincipalPlane(offset=0.0,
principalPlane=XYPLANE)
mdb.models[model_name].parts[part_name].DatumPlaneByPrincipalPlane(offset=0.0,
principalPlane=YZPLANE)
mdb.models[model_name].parts[part_name].DatumPlaneByPrincipalPlane(offset=0.0,
principalPlane=XZPLANE)
mdb.models[model_name].parts[part_name].PartitionCellByDatumPlane(cells=
mdb.models[model_name].parts[part_name].cells.getSequenceFromMask(('[#1 ]',
), ), datumPlane=mdb.models[model_name].parts[part_name].datums[2])
mdb.models[model_name].parts[part_name].PartitionCellByDatumPlane(cells=
mdb.models[model_name].parts[part_name].cells.getSequenceFromMask(('[#3 ]',
), ), datumPlane=mdb.models[model_name].parts[part_name].datums[3])

72
mdb.models[model_name].parts[part_name].PartitionCellByDatumPlane(cells=
mdb.models[model_name].parts[part_name].cells.getSequenceFromMask(('[#f ]',
), ), datumPlane=mdb.models[model_name].parts[part_name].datums[4])
mdb.models[model_name].parts[part_name].seedPart(deviationFactor=0.1,
minSizeFactor=0.1, size=mesh_size)
mdb.models[model_name].parts[part_name].generateMesh()

# material
mdb.models[model_name].Material(name=material_name)
mdb.models[model_name].materials[material_name].Density(table=((1e-15, ), ))
mdb.models[model_name].materials[material_name].Elastic(table=((10000.0, 0.435),
))
mdb.models[model_name].materials[material_name].Plastic(table=((5.0, 0.0), (
10.43214307, 0.05), (11.92359568, 0.1), (12.97927878, 0.15), (13.82454245,
0.2), (14.5413692, 0.25), (15.17007457, 0.3), (15.7338498, 0.35), (
16.24741436, 0.4), (16.72076823, 0.45), (17.16105352, 0.5), (17.57357103,
0.55), (17.96237662, 0.6), (18.33065144, 0.65), (18.68094233, 0.7), (
19.01532395, 0.75), (19.33551151, 0.8), (19.64294133, 0.85), (19.9388297,
0.9), (20.2242168, 0.95), (20.5, 1.0)))
mdb.models[model_name].HomogeneousSolidSection(material=material_name, name=
'Section-1', thickness=None)
mdb.models[model_name].parts[part_name].Set(cells=
mdb.models[model_name].parts[part_name].cells.getSequenceFromMask(('[#ff ]',
), ), name='Set-1')
mdb.models[model_name].parts[part_name].SectionAssignment(offset=0.0,
offsetField='', offsetType=MIDDLE_SURFACE, region=
mdb.models[model_name].parts[part_name].sets['Set-1'], sectionName=
'Section-1', thicknessAssignment=FROM_SECTION)

# loop to copy parts


for i in range(np_total-1):
p_name = part_org_name + repr(i+2)
mdb.models[model_name].Part(name=p_name, objectToCopy=
mdb.models[model_name].parts[part_name])

# loop to read and assign particle positions


mdb.models[model_name].rootAssembly.DatumCsysByDefault(CARTESIAN)

73
# define particle path and name
f = open('D:\YuZhang\p_pos.txt', 'r')
for i in range(np_total):
p_name = part_org_name + repr(i+1)
p_name2 = p_name + "-1"
mdb.models[model_name].rootAssembly.Instance(dependent=ON, name=p_name2
, part=mdb.models[model_name].parts[p_name])

content = f.readline();
mdb.models[model_name].rootAssembly.translate(instanceList=(p_name2, ),
vector= [float(content) for content in content.split()])
f.close()

# Lower Boundary
mdb.models['Model-1'].ConstrainedSketch(name='__profile__', sheetSize=200.0)
mdb.models['Model-1'].sketches['__profile__'].ConstructionLine(point1=(0.0,
-100.0), point2=(0.0, 100.0))
mdb.models['Model-1'].sketches['__profile__'].FixedConstraint(entity=
mdb.models['Model-1'].sketches['__profile__'].geometry[2])
mdb.models['Model-1'].sketches['__profile__'].Line(point1=(0.0, 0.0), point2=(
25, 0.0))
mdb.models['Model-1'].sketches['__profile__'].HorizontalConstraint(
addUndoState=False, entity=
mdb.models['Model-1'].sketches['__profile__'].geometry[3])
mdb.models['Model-1'].sketches['__profile__'].PerpendicularConstraint(
addUndoState=False, entity1=
mdb.models['Model-1'].sketches['__profile__'].geometry[2], entity2=
mdb.models['Model-1'].sketches['__profile__'].geometry[3])
mdb.models['Model-1'].sketches['__profile__'].CoincidentConstraint(
addUndoState=False, entity1=
mdb.models['Model-1'].sketches['__profile__'].vertices[0], entity2=
mdb.models['Model-1'].sketches['__profile__'].geometry[2])
mdb.models['Model-1'].Part(dimensionality=THREE_D, name='lower B', type=
ANALYTIC_RIGID_SURFACE)
mdb.models['Model-1'].parts['lower B'].AnalyticRigidSurfRevolve(sketch=
mdb.models['Model-1'].sketches['__profile__'])
del mdb.models['Model-1'].sketches['__profile__']

74
# Upper Boundary
mdb.models['Model-1'].ConstrainedSketch(name='__profile__', sheetSize=200.0)
mdb.models['Model-1'].sketches['__profile__'].ConstructionLine(point1=(0.0,
-100.0), point2=(0.0, 100.0))
mdb.models['Model-1'].sketches['__profile__'].FixedConstraint(entity=
mdb.models['Model-1'].sketches['__profile__'].geometry[2])
mdb.models['Model-1'].sketches['__profile__'].Line(point1=(0.0, 10.0), point2=(
25, 10.0))
mdb.models['Model-1'].sketches['__profile__'].HorizontalConstraint(
addUndoState=False, entity=
mdb.models['Model-1'].sketches['__profile__'].geometry[3])
mdb.models['Model-1'].sketches['__profile__'].PerpendicularConstraint(
addUndoState=False, entity1=
mdb.models['Model-1'].sketches['__profile__'].geometry[2], entity2=
mdb.models['Model-1'].sketches['__profile__'].geometry[3])
mdb.models['Model-1'].sketches['__profile__'].CoincidentConstraint(
addUndoState=False, entity1=
mdb.models['Model-1'].sketches['__profile__'].vertices[0], entity2=
mdb.models['Model-1'].sketches['__profile__'].geometry[2])
mdb.models['Model-1'].Part(dimensionality=THREE_D, name='upper B', type=
ANALYTIC_RIGID_SURFACE)
mdb.models['Model-1'].parts['upper B'].AnalyticRigidSurfRevolve(sketch=
mdb.models['Model-1'].sketches['__profile__'])
del mdb.models['Model-1'].sketches['__profile__']

# Die wall
mdb.models['Model-1'].ConstrainedSketch(name='__profile__', sheetSize=200.0)
mdb.models['Model-1'].sketches['__profile__'].ArcByCenterEnds(center=(0.0, 0.0)
, direction=CLOCKWISE, point1=(25, 0.0), point2=(0.0, -25))
mdb.models['Model-1'].sketches['__profile__'].ArcByCenterEnds(center=(0.0, 0.0)
, direction=CLOCKWISE, point1=(0.0, -25), point2=(-25, 0.0))
mdb.models['Model-1'].sketches['__profile__'].ArcByCenterEnds(center=(0.0, 0.0)
, direction=CLOCKWISE, point1=(-25, 0.0), point2=(0.0, 25))
mdb.models['Model-1'].sketches['__profile__'].ArcByCenterEnds(center=(0.0, 0.0)
, direction=CLOCKWISE, point1=(0.0, 25), point2=(25, 0.0))
mdb.models['Model-1'].Part(dimensionality=THREE_D, name='die wall', type=
ANALYTIC_RIGID_SURFACE)
mdb.models['Model-1'].parts['die wall'].AnalyticRigidSurfExtrude(depth=50,

75
sketch=mdb.models['Model-1'].sketches['__profile__'])
del mdb.models['Model-1'].sketches['__profile__']
mdb.models['Model-1'].rootAssembly.Instance(dependent=ON, name='die wall-1',
part=mdb.models['Model-1'].parts['die wall'])
mdb.models['Model-1'].rootAssembly.Instance(dependent=ON, name='lower B-1',
part=mdb.models['Model-1'].parts['lower B'])
mdb.models['Model-1'].rootAssembly.Instance(dependent=ON, name='upper B-1',
part=mdb.models['Model-1'].parts['upper B'])

# Step
mdb.models['Model-1'].ExplicitDynamicsStep(massScaling=((SEMI_AUTOMATIC, MODEL,
AT_BEGINNING, 0.0, time_step, BELOW_MIN, 0, 0, 0.0, 0.0, 0, None), ), name=
'Step-1', previous='Initial', timePeriod=Time_period)

# Reference Point
mdb.models['Model-1'].parts['lower B'].ReferencePoint(point=
mdb.models['Model-1'].parts['lower B'].vertices[1])
mdb.models['Model-1'].parts['upper B'].ReferencePoint(point=
mdb.models['Model-1'].parts['upper B'].vertices[1])
mdb.models['Model-1'].parts['die wall'].ReferencePoint(point=
mdb.models['Model-1'].parts['die wall'].vertices[4])

# History outpot
mdb.models['Model-1'].rootAssembly.regenerate()
mdb.models['Model-1'].rootAssembly.Set(name='Set-1', referencePoints=(
mdb.models['Model-1'].rootAssembly.instances['upper B-1'].referencePoints[2],
))
mdb.models['Model-1'].HistoryOutputRequest(createStepName='Step-1', name=
'H-Output-2', rebar=EXCLUDE, region=
mdb.models['Model-1'].rootAssembly.sets['Set-1'], sectionPoints=DEFAULT,
variables=('U1', 'U2', 'U3', 'UR1', 'UR2', 'UR3', 'UT', 'RF1', 'RF2',
'RF3', 'RM1', 'RM2', 'RM3', 'RT'))

# Constraints
mdb.models['Model-1'].rootAssembly.Surface(name='Surf-1', side2Faces=
mdb.models['Model-1'].rootAssembly.instances['upper B-1'].faces.getSequenceFromMask(
('[#1 ]', ), ))
mdb.models['Model-1'].RigidBody(name='Constraint-1', refPointRegion=Region(

76
referencePoints=(
mdb.models['Model-1'].rootAssembly.instances['upper B-1'].referencePoints[2],
)), surfaceRegion=mdb.models['Model-1'].rootAssembly.surfaces['Surf-1'])
mdb.models['Model-1'].rootAssembly.Surface(name='Surf-2', side1Faces=
mdb.models['Model-1'].rootAssembly.instances['die wall-1'].faces.getSequenceFromMask(
('[#1 ]', ), ))
mdb.models['Model-1'].RigidBody(name='Constraint-2', refPointRegion=Region(
referencePoints=(
mdb.models['Model-1'].rootAssembly.instances['die wall-1'].referencePoints[2],
)), surfaceRegion=mdb.models['Model-1'].rootAssembly.surfaces['Surf-2'])
mdb.models['Model-1'].rootAssembly.Surface(name='Surf-3', side1Faces=
mdb.models['Model-1'].rootAssembly.instances['lower B-1'].faces.getSequenceFromMask(
('[#1 ]', ), ))
mdb.models['Model-1'].RigidBody(name='Constraint-3', refPointRegion=Region(
referencePoints=(
mdb.models['Model-1'].rootAssembly.instances['lower B-1'].referencePoints[2],
)), surfaceRegion=mdb.models['Model-1'].rootAssembly.surfaces['Surf-3'])

# Interaction
mdb.models['Model-1'].ContactProperty('IntProp-1')
mdb.models['Model-1'].interactionProperties['IntProp-1'].TangentialBehavior(
dependencies=0, directionality=ISOTROPIC, elasticSlipStiffness=None,
formulation=PENALTY, fraction= friction_coefficient, maximumElasticSlip=FRACTION,
pressureDependency=OFF, shearStressLimit=None, slipRateDependency=OFF,
table=((0.05, ), ), temperatureDependency=OFF)
mdb.models['Model-1'].ContactExp(createStepName='Step-1', name='Int-1')
mdb.models['Model-1'].interactions['Int-1'].includedPairs.setValuesInStep(
stepName='Step-1', useAllstar=ON)
mdb.models['Model-1'].interactions['Int-1'].contactPropertyAssignments.appendInStep(
assignments=((GLOBAL, SELF, 'IntProp-1'), ), stepName='Step-1')

# BCs
mdb.models['Model-1'].rootAssembly.Set(name='Set-5', referencePoints=(
mdb.models['Model-1'].rootAssembly.instances['lower B-1'].referencePoints[2],
))
mdb.models['Model-1'].DisplacementBC(amplitude=UNSET, createStepName='Step-1',
distributionType=UNIFORM, fieldName='', fixed=OFF, localCsys=None, name=
'BC-1', region=mdb.models['Model-1'].rootAssembly.sets['Set-5'], u1=0.0,

77
u2=0.0, u3=0.0, ur1=0.0, ur2=0.0, ur3=0.0)
del mdb.models['Model-1'].boundaryConditions['BC-1']
mdb.models['Model-1'].rootAssembly.Set(name='Set-6', referencePoints=(
mdb.models['Model-1'].rootAssembly.instances['lower B-1'].referencePoints[2],
))
mdb.models['Model-1'].DisplacementBC(amplitude=UNSET, createStepName='Step-1',
distributionType=UNIFORM, fieldName='', fixed=OFF, localCsys=None, name=
'lower B', region=mdb.models['Model-1'].rootAssembly.sets['Set-6'], u1=0.0,
u2=0.0, u3=0.0, ur1=0.0, ur2=0.0, ur3=0.0)
mdb.models['Model-1'].rootAssembly.Set(name='Set-7', referencePoints=(
mdb.models['Model-1'].rootAssembly.instances['die wall-1'].referencePoints[2],
))
mdb.models['Model-1'].DisplacementBC(amplitude=UNSET, createStepName='Step-1',
distributionType=UNIFORM, fieldName='', fixed=OFF, localCsys=None, name=
'die wall', region=mdb.models['Model-1'].rootAssembly.sets['Set-7'], u1=0.0
, u2=0.0, u3=0.0, ur1=0.0, ur2=0.0, ur3=0.0)
mdb.models['Model-1'].rootAssembly.Set(name='Set-8', referencePoints=(
mdb.models['Model-1'].rootAssembly.instances['upper B-1'].referencePoints[2],
))
mdb.models['Model-1'].VelocityBC(amplitude=UNSET, createStepName='Step-1',
distributionType=UNIFORM, fieldName='', localCsys=None, name='upper B',
region=mdb.models['Model-1'].rootAssembly.sets['Set-8'], v1=0.0, v2= Boundary_velocity,
v3=0.0, vr1=0.0, vr2=0.0, vr3=0.0)

# Job
mdb.Job(activateLoadBalancing=False, atTime=None, contactPrint=OFF,
description='', echoPrint=OFF, explicitPrecision=SINGLE, historyPrint=OFF,
memory=90, memoryUnits=PERCENTAGE, model='Model-1', modelPrint=OFF,
multiprocessingMode=DEFAULT, name='4particles', nodalOutputPrecision=SINGLE
, numCpus=7, numDomains=7, parallelizationMethodExplicit=DOMAIN, queue=None
, resultsFormat=ODB, scratch='', type=ANALYSIS, userSubroutine='',
waitHours=0, waitMinutes=0)
mdb.models['Model-1'].rootAssembly.Set(name='Set-9', nodes=
mdb.models['Model-1'].rootAssembly.instances['Sphere-1-1'].nodes.getSequenceFromMask(
mask=('[#2 ]', ), )+\
mdb.models['Model-1'].rootAssembly.instances['Sphere-2-1'].nodes.getSequenceFromMask(
mask=('[#2 ]', ), )+\
mdb.models['Model-1'].rootAssembly.instances['Sphere-3-1'].nodes.getSequenceFromMask(

78
mask=('[#2 ]', ), )+\
mdb.models['Model-1'].rootAssembly.instances['Sphere-4-1'].nodes.getSequenceFromMask(
mask=('[#2 ]', ), )+\
mdb.models['Model-1'].rootAssembly.instances['Sphere-5-1'].nodes.getSequenceFromMask(
mask=('[#2 ]', ), )+\
mdb.models['Model-1'].rootAssembly.instances['Sphere-6-1'].nodes.getSequenceFromMask(
mask=('[#2 ]', ), )+\
mdb.models['Model-1'].rootAssembly.instances['Sphere-7-1'].nodes.getSequenceFromMask(
mask=('[#2 ]', ), )+\
mdb.models['Model-1'].rootAssembly.instances['Sphere-8-1'].nodes.getSequenceFromMask(
mask=('[#2 ]', ), )+\
mdb.models['Model-1'].rootAssembly.instances['Sphere-9-1'].nodes.getSequenceFromMask(
mask=('[#2 ]', ), ))
mdb.models['Model-1'].FieldOutputRequest(createStepName='Step-1', name=
'F-Output-2', rebar=EXCLUDE, region=
mdb.models['Model-1'].rootAssembly.sets['Set-9'], sectionPoints=DEFAULT,
variables=('COORD', ))

79
Appendix B. Python script to of probability density calculation
####calculate euclidean distance between two 3d points
#import numpy as np
import math

##import file and read data by each line


file_name = "den_055"
coord_file_name = file_name + '_coord.txt'
overlap_file_name = file_name + '_overlap.txt'
angle_file_name = file_name + '_angle.txt'
cn_file_name = file_name + '_cn.txt'
contact_number_name = file_name + '_number.txt'
##store reading data from lines
nptotal = 0
coord_file = open(coord_file_name)
overlap_file = open(overlap_file_name, 'w+')
angle_file = open(angle_file_name, 'w+')
cn_file = open(cn_file_name, 'w+')
number_file = open(contact_number_name , 'w+')
pPos = []
for line in coord_file:
if(line.find('Zone SolutionTime')>=0):
continue
else:
vector= [float(line) for line in line.split()]
pPos.append(vector)
nptotal += 1

# get number of contact, mininum and maximum overlap


nc = 0 #initialise contact number
overlap_min = 1
overlap_max = 0
overlap = []
angle = []

# coordination number list


cnlist = []
for i in range(nptotal) :
80
cnlist.append(0)

for i in range(nptotal) :
for j in range(i+1,nptotal):

# 2 particle positions and diameter


x1 = pPos[i][0]
y1 = pPos[i][1]
z1 = pPos[i][2]
r1 = pPos[i][3]

x2 = pPos[j][0]
y2 = pPos[j][1]
z2 = pPos[j][2]
r2 = pPos[j][3]

xy = ((x1-x2)**2+(y1-y2)**2)**0.5
zz = abs(z1-z2)

ijPos = (xy**2+zz**2)**0.5

ijoverlap = ((r1+r2)/2)*1.05 - ijPos


if(ijoverlap > 0.):
nc += 1
cnlist[i] +=1
cnlist[j] +=1

ijoverlap /=(r1+r2)/2
overlap.append(ijoverlap)
overlap_max = max(ijoverlap,overlap_max)
overlap_min = min(ijoverlap,overlap_min)

# vertical has zero angle


if(xy == 0.):
ijAng = 90
angle.append(ijAng)
else:

81
ijAng = math.atan(zz/xy) * 180/math.pi
angle.append(ijAng)

# get overlap distribution


nbin = 15
dbin = (overlap_max - overlap_min)/nbin
bin = []
for ibin in range (nbin):
bin.append(0)

for i in range(nc):
ibin = int(overlap[i]/dbin)
if (ibin < nbin):
bin[ibin] += 1

for ibin in range(nbin):


overlap_file.write(str((ibin+0.5)*dbin) + " "+ str(bin[ibin]/(nc*dbin))+"\n")

# get angle distribution


ngbin = 19
dgbin = 5
gbin = []
for igbin in range (ngbin):
gbin.append(0)

for i in range(nc):
igbin = int(angle[i]/dgbin)
if (igbin < ngbin):
gbin[igbin] += 1

for igbin in range(ngbin):


angle_file.write(str((igbin)*dgbin) + " "+ str(gbin[igbin]/(nc*dgbin))+"\n")

for i in range(nptotal):
cn_file.write(str(cnlist[i])+"\n")
##Count
from collections import Counter
count = Counter(cnlist)

82
number_file.write(str(count))
# close files
coord_file.close()
overlap_file.close()
angle_file.close()
cn_file.close()
number_file.close()
print(nc, overlap_max, overlap_min)

83

You might also like