Sei sulla pagina 1di 58

Application of user defined

subroutine UMESHMOTION in
ABAQUS for simulating dry
rolling/sliding wear

Masters Thesis of
Basavaraj Kanavalli
Master of Science (Scientific Computing)
Royal Institute of Technology (KTH),
Stockholm, Sweden.
Supervised by:
Dr. Vishwanath Hegadekatte
Institute for Reliability of Components and Systems
University of Karlsruhe, Germany.
Prof. Sren Andersson
Department of Machine Design,
Royal Institute of Technology (KTH), Stockholm, Sweden.

Abstract

Twin-disc rolling/sliding tribometer experiments try to mimic the rolling/sliding


contact experienced by micro-machines, e.g., between the teeth of two mating
micro-gears. In this work, a user defined subroutine UMESHMOTION, in the
commercial finite element code, ABAQUS, has been applied to simulate wear in a
twin-disc tribometer experiments, conducted for defined slips. ABAQUS invokes
the adaptive meshing algorithm after the convergence of the equilibrium equations of the contact problem, which further, invokes the user defined sub-routine
UMESHMOTION. UMESHMOTION is coded to compute the local wear using
the generalized Archards wear model and it integrates the wear depth over the
sliding distance using Euler integration scheme. In the absence of wear coefficient
data from such experiments, it is assumed to be the same as identified from pinon-disc experiments (Herz et al., 2004; J. Schneider & Herz, 2005). The computed
wear is applied on each surface node as mesh-constraint by the adaptive meshing
algorithm. The resulting equilibrium loss is corrected by solving the last time
increment. Thus the geometry and pressures are updated. Simulations were carried out with different applied loads (with constant slip) and different slips (with
constant load) and the results obtained, are presented. The results obtained from
UMESHMOTION are discussed by comparing them with the Global Incremental
Wear Model (GIWM) and the Wear-Processor (Hegadekatte, 2006a).

iii

Acknowledgement

Success of any work would not be completed unless we mention the names of
the people who contributed in the work directly or indirectly. While I cannot
even begin to thank all those who have unknowingly helped me throughout my
studies, I would like to express my deep felt gratitude to some people who have
helped me with useful suggestions and guidance.
First of all, I would like to express my deep felt gratitude for my adviser, Dr.
Vishwanath Hegadekatte, without whose advises, encouragement and patience,
this work would not have been possible, even to think of. The suggestions, tips
and advises given by Vishwanath were immensly helpful for my thesis. I would like
to thank Professor Oliver Kraft, Forschungszentrum Karlsruhe (FZK), for giving
me this opportunity to work on my master thesis in an international research
atmosphere. I would like to express my deep felt gratitude for Professor Norbert
Huber for his advises and support during the course of this work. I would also
like to thank Yixiang Gan, at IMF II, FZK, for his useful tips and suggestions.
I would like to express my deep felt gratitude for Swedish Government, Dr Lennart
Edsberg and Dr Katarina Gustavsson, for keeping faith in me and giving me
an opportunity to study at Royal Institute of Technology (KTH). I am greatly
indebted for the scholarship awarded by STINT for my studies at KTH. I would
like to express my sincere thanks to Professor Sren Andersson for accepting me
as his student and guiding me, from Sweden, during my master thesis, amidst
his busy schedule. I would also like to thank DAAD and IAESTE for helping
me in getting my visa for my thesis at FZK, Germany. Especially, I am greatly
indebted to the cooperation that I received from Barhammar Corolla, Jasmine
Agassi, Gerd Aye and Maike Weisskopf.
v

Last but not the least, I want to acknowledge my family members for their continuous support and encouragement. This work is dedicated to my mother Mrs.
Kamaladevi Kanavalli and my father Mr. Chudamani Kanavalli.
Basavaraj Kanavalli
Karlsruhe, 13 June 2006

vi

Contents
Abstract

iii

Acknowledgement

1 Introduction

1.1
1.2

Tribometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Overview of chapters . . . . . . . . . . . . . . . . . . . . . . . . .

2 Wear Simulation of Twin Disc Rolling/Sliding Tribometer


2.1
2.2

Formulation of contact problem using FEM . . . . . . . . . . . .


Wear simulation strategy . . . . . . . . . . . . . . . . . . . . . . .
2.2.1

2.3

2
3
5
5
8

Wear model for twin-disc rolling/sliding tribometer . . . .


2.2.1.1 Generalized Archards wear model . . . . . . . .

11
12

Wear simulation of twin-disc rolling/sliding tribometer. . . . . . .


2.3.1 Adaptive meshing . . . . . . . . . . . . . . . . . . . . . . .

14
15

2.3.1.1
2.3.1.2
2.3.2

Mesh sweeping . . . . . . . . . . . . . . . . . . .
Advection . . . . . . . . . . . . . . . . . . . . . .

16
16

Wear-Simulator . . . . . . . . . . . . . . . . . . . . . . . .
2.3.2.1 Computation of inward surface normals . . . . .

17
19

3 Wear Simulation Results

23

3.1
3.2

FE model of the twin-disc rolling/sliding tribometer . . . . . . . .


Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

23
26

3.3

Wear simulation with different loads and slips . . . . . . . . . . .


3.3.1 Wear Simulations with different loads . . . . . . . . . . . .

28
28

3.3.2 Wear Simulations with different slips . . . . . . . . . . . .


Comparison with GIWM . . . . . . . . . . . . . . . . . . . . . . .

29
30

3.4

vii

4 Discussion and conclusion


4.1 Numerical integration method . . . . . . . . . . . . . . . . . . . .
4.2 Pressure field updated by ABAQUS . . . . . . . . . . . . . . . . .
4.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

33
35
38
41

5 Summary

43

List of Figures

45

List of Tables

46

Bibliography

49

viii

Chapter 1

Introduction

Micro-machines are part of a rapidly emerging technology, finding wide variety of


applications in our day today life. The reliability of micro-machines is, therefore,
a critical factor. The primary factors determining the performance and reliability
of micro-machines are adhesion, friction, wear, film stress, fracture strength and
fatigue (Romig et al., 2003). Out of these modes of failure, excessive wear is the
most critical parameter, reducing the life span of micro-machines substantially,
owing to their high operating speeds and high surface to volume ratio. For
example, Williams (2001) observed that 1 minute of life for a micro-machine
represents a degree of wear and degradation equal to more than 10 years as in a
well designed watch bearing. In Figure 1.1 (a) shows a linear rack micro-motor.
We can see that, between the two mating gear teeth, there will be rolling/sliding
contact and between the gear and shaft, on which it is mounted, there will be
sliding contact. Figure 1.1 (b) shows how a gap is formed between the center of
the micro gear and the shaft, because of wear. Compared to the research work
done on modeling and simulations of wear in the macro world, less effort has
been made to study wear in micro-machines, due to obvious difficulties at micro
dimensions. For example, one of the difficulties is that the surface tension of
fluids used in the final rinse can glue the parts of a micro-machine together.

1.1. Tribometers
(a)

(b)

Figure 1.1: (a) Picture showing linear rack micro-motor, (b) Picture showing gap
formation on the gear, because of wear. Courtesy: Dannelle M. Tanner (Sandia
National Laboratories).

1.1

Tribometers

At present, in-situ wear measurements are the most realistic methods to predict
useful life-span of micro-machines. But manufacturing of prototypes and measuring of wear to estimate the useful life span of micro-machines involves extremely
high costs, both in terms of money and time. Because of this, wear rate is usually
determined experimentally using a tribometer like the twin-disc rolling/sliding
tribometer (see Figure 1.2). These tribometers try to mimic the contact conditions experienced by micro-machines. e. g., the twin-disc rolling/sliding tribometer tries to mimic the rolling/sliding contact experienced by teeth of two
mating micro-gears. The specimens used in tribometers are manufactured exactly
by the same production/manufacturing processes as that of the micro-machines
and therefore are assumed to posses same microstructure as that of the micromachines. In twin-disc rolling/sliding tribometer, the two discs are loaded against
each other and are rotated at different speeds. The weight loss from either of the
discs, is measured after certain number of rotations. Wear is expressed as worn
volume per unit load and per unit sliding distance which is called as the dimensional wear coefficient, kD . Dimensional wear coefficient is the most important
wear parameter determined from a tribometer.
Tribometer gives a qualitative picture of the suitability of a material combination

1. Introduction

Figure 1.2: Twin disc rolling/sliding tribometer.


for a given application and does not predict the useful life span of micro-machines.
A wear simulation strategy that uses the finite element method (FEM), has been
proposed by Hegadekatte et al. (2005) and the working of the wear simulation
tool, Wear-Processor is explained in brief, in the second chapter.

1.2

Overview of chapters

In this chapter we discussed about the importance of wear in micro-machines and


the role of tribometers in wear measurement. In chapter two, formulation of contact problem using FEM will be discussed. Later in this chapter wear simulation
strategy adopted by Hegadekatte et al. (2005) for simulating dry sliding wear
in pin-on-disc tribometer is discussed. This will be followed by discussion about
modification of Archards wear model to simulate wear in twin-disc rolling/sliding
tribometer. In the same chapter, adaptive meshing and working of UMESHMOTION, which is a user defined subroutine in commercial FE package, ABAQUS,
to simulate dry sliding wear in twin-disc rolling/sliding tribometer, is explained
in detail. In chapter three, the finite element model used in this work is presented. In this chapter, the results obtained from the FE based wear simulation
using UMESHMOTION are presented. in chapter four, the results obtained from
UMESHMOTION are compared with two different simulation tools, namely, the

1.2. Overview of chapters

Global Incremental Wear Model (GIWM) and the Wear-Processor. Finally, a


summary of this work is given in chapter five.

Chapter 2
Wear Simulation of Twin Disc
Rolling/Sliding Tribometer

The whole process of simulating wear on any general tribosystem consists of two
main tasks, namely, solution of the contact problem and using this solution to
compute wear. In the first part of this chapter the formulation of contact problem
using FEM is explained in brief. In the second part, a brief description about
the wear simulation strategy adopted by Hegadekatte et al. (2005), for simulating wear in pin-on-disc tribometer and modification of the Archards wear
model to compute wear in twin-disc tribometer is given. In the third part, detailed description of adoptive meshing and the working of user defined subroutine
UMESHMOTION and its application for wear simulation is given.

2.1

Formulation of contact problem using FEM

In FE analysis, the contact conditions are applied as constraints. Unlike boundary conditions which remain fixed throughout the analysis, the constraints are
discontinuous. Two methods for applying constraints are very popular, namely
the Lagrange multiplier method, and the penalty method. In the following lines,
the formulation of the contact problem in FEM using the Lagrange multiplier

2.1. Formulation of contact problem using FEM

method for application of contact constraints will be discussed in brief (Bathe,


1996). The assumptions made are:
1. arbitrarily shaped body,
2. small deformation,
3. linear elastic material property,
4. static problem and
5. external forces include volume forces, surface forces and concentrated forces
(which include the contact forces).
Assuming a linear elastic continuum, the total potential is given by:
Z
1
T DdV
(u, ) =
2 V
|
{z
}
Strain Energy
Z

uT f V dV
| V {z
}
Work done by body forces
Z
T
uS f S dS
| S {z
}
Work done by surface forces
X T

uC FC
|C {z }
Work done by concentrated forces
X T
T
+
k (Qk uk k ),

(2.1)

{z
}
|k
Work done by contact forces
(Lagrange multiplier method),
where, is the strain tensor, D is the stress-strain matrix for the linear elastic
material, u is the displacement vector, f V is the body force vector, f S is the vector
T
of surface forces, FC is the vector of concentrated forces, k is the vector of

2. Wear Simulation of Twin Disc Rolling/Sliding Tribometer

contact forces for the contact node k, Qk is the contact matrix for contact node
k, k is the penetration of contact node k, V is the volume, S is the surface, ()C is
the superscript for concentrated forces and the respective displacements and ()k
is the superscript for contact node k.
By applying the principle of minimum potential energy, i.e, = 0, we can
formulate the governing equations for the body.
Z
Z
Z
T
T
T V
=
DdV
u f dV uS f S dS
V

X
T
T
T
T
T
uC f C +
[k (Qk uk k ) + k Qk uk ].

(2.2)

In finite element analysis the domain is approximated as a set of discrete subdomains called as finite elements with elements being interconnected at nodal
points on the element boundaries. Therefore the equilibrium equations that correspond to the nodal point displacements of the assemblage of finite elements is
written as a sum of integrations over the volume and the areas of all finite elements. Then applying the principle of stationarity or the principle of minimum
potential energy, we get the following governing equations for the system, (for
more details refer Hegadekatte (2006b)),

T
U

"
XZ
M

#
T

B(M ) D(M ) B(M ) dV (M ) U


V

T
= + U

(M )

XZ
M

T
+ U

HB(M ) f B(M ) dV (M )
V (M )

XZ

HS(M ) f S(M ) dS (M )

S (M )

T
U

k(M )

T
+ U
Q
h

FC

C (M )

i
T
T
k

Q U ,

(2.3)

k(M )

where the superscript M stands for elements in the domain, B(M ) is the strain
is the vector
displacement matrix, H(M ) is displacement interpolation matrix, U
of global displacements at all nodal points. F is the vector of all the concentrated

2.2. Wear simulation strategy

is the vector of all the contact forces, Q is the complete contact matrix
forces,
of the assemblage and is the vector for all penetrations. Since, for a linear
elastic continuum, the principle of minimum potential energy is identical to the
principle of virtual displacement, the unknown nodal point displacement can be
obtained from Equation (2.3) by imposing unit virtual displacement in turn at
= I.
all displacement components. In this way we have U
The equilibrium equations of the element assemblage, corresponding to the nodal
point displacements are given by:
"
#"
# "
# "
#

K Q
U
R
Rk
=
+
,
(2.4)

QT 0

where,
K=

P R
M

V (M )

B(M ) D(M ) B(M ) dV (M )

R = RB + RS
=

P R
M

V (M )

HB

(M )T

fB

(M )

dV (M ) +

P R
M

S (M )

HS(M ) f S(M ) dS (M )

T .
Rk = Q
For nonlinear problems the system of equations is formulated in an appropriate
way and is solved using the N ewton Raphson M ethod (Deuflhard, 2004; Ypma,
1995).

2.2

Wear simulation strategy

In this section the wear simulation strategy adopted by Hegadekatte et al. (2005)
for simulating dry sliding wear in pin-on-disc tribometer will be explained, in
brief. A wear simulation tool, Wear-Processor, has been developed, which works
in conjunction with a commercial FE package ABAQUS. In the following lines the
working of Wear-Processor, to simulate wear on the pin and the disc, is explained
in brief. The contact simulation of the FE model is solved using the commercial
FE package, ABAQUS. The contact pressures on the surface nodes are calculated
using the stress tensors and surface normals and are used to compute wear using

2. Wear Simulation of Twin Disc Rolling/Sliding Tribometer

the Archards wear law. Archards wear law is the simplest wear model available
in the literature and can be expressed as,

V
s

FN
H
= kD FN ,
= k

(2.5)

where V is the volume of material removed, s is the sliding distance, FN is the


applied normal load, k is the dimensionless wear coefficient and kD [mm3 /N mm]
is the dimensional wear coefficient. The hardness (ratio of load over projected
area), H, is that of the softer material. By dividing both sides of Equation (2.5)
by real area of contact we get the expression for wear depth,
h
= kD p,
s

(2.6)

where p is the pressure. This wear model is applicable on the global scale and
in this work it is assumed that it is also applicable on the local scale. That
is, contact pressures at each node on the surface can be used to compute wear
on these nodes. Therefore, Equation (2.6) is discretized with respect to sliding
distance as,
dh
= kD p.
ds

(2.7)

As the disc wears out, the contact area increases and hence the contact pressure
decreases. Therefore, Equation (2.7) is integrated over the sliding distance by
using Euler integration scheme, to obtain expression for wear depth, as shown
below,
hi+1 = hi + kD pi ds,

(2.8)

where hi+1 is the total wear up to the (i+1)th wear increment, and hi is total wear
up to the ith wear increment, pi is the contact pressure during ith wear increment
and kD pi ds is the wear depth for the (i + 1)th (current) wear increment. The
Equation (2.8) computes wear on a node which is in continuous contact with
its mating surface. e.g., for a node on the surface of the pin. In pin-on-disc
tribometer, any point on the disc comes in contact with the pin surface, only
once in one complete revolution of the pin over disc. It is during this period, that

10

2.2. Wear simulation strategy

the wear takes place on each of these disc surface nodes, depending on the contact
pressure that the considered node experiences. As the pin moves over the disc,
the contact pressure on the considered node, increases from zero to maximum and
then decreases to zero. Therefore, to compute wear on the disc surface nodes,
contact pressure is integrated over the circumferential length at a given radius,
R, of the disc, as,
Z

hi+1 = hi + kD

pi Rd,

(2.9)

=0

where is the circumferential coordinate. In order to compute the integral in


Equation (2.9) it is necessary that the finite element surface nodes are located
exactly along the circumference of the various streamlines on the disc. This
restriction on the mesh is over come by assuming an imaginary grid of points on
the disc such that these grid points are located at definite pre-determined radii
of the circumferential streamlines on the disc. The grid points are first identified
with the three dimensional surface element faces that they are located on. The
contact pressures at these grid points are obtained by interpolating the contact
pressures calculated at the surface nodes. With the help of the computed contact
pressures at grid points, the integral in the Equation (2.9) is calculated. Then,
wear depth is computed using Equation (2.9). This wear depth is applied to all
surface nodeson the pertcular streamline, in the inward normal direction. If the
wear depth after certain amount of sliding is more than the product of surface
element height and a pre-set factor, then the domain is re-meshed. The process
is repeated until the desired amount of sliding distance is reached.
This work is based on the wear simulation strategy adopted by Hegadekatte
et al. (2005), but for twin-disc rolling/sliding tribometer. The computation of
wear on the disc in Hegadekatte et al. (2005) is similar to the computation of
wear in twin-disc rolling/sliding tribometer. In both the cases, the point under
consideration comes in contact with its mating surface once per rotation of the
disc. The considered point enters the contact interface and then leaves the contact
interface. Therefore, Equation (2.9) is modified suitably to compute wear in twindisc rolling/sliding tribometer. In the next sub-section the modification of the
Equation (2.9) to compute wear in twin-disc rolling/sliding tribometer along with
the assumptions made in this work are explained in detail.

2. Wear Simulation of Twin Disc Rolling/Sliding Tribometer


(a)

11

(b)

Figure 2.1: (a) Schematic view of the twin-disc tribometer; (b) Figure showing
the contact simulation geometric model for twin-disc tribometer.

2.2.1

Wear model for twin-disc rolling/sliding tribometer

When two circular rotating bodies come in contact with each other and they have
same tangential velocity at all points of contact, then they are said to exhibit
pure rolling contact. In pure rolling there will be no slip at the contact interface.
However in reality it is very difficult to find a contact situation which exhibits
pure rolling. Local sliding takes place in a part of the contact. Most of the rolling
contacts are, therefore, actually rolling and sliding contacts (Spiegelberg, 2005).
Therefore, in this work, twin-disc rolling/sliding tribometer with definite slip,
between the two discs, is considered for wear simulation and also the experiments
are conducted by (Herz et al., 2004; J. Schneider & Herz, 2005) for the same
reason. The existence of slip between the discs together with normal load acting
on the them, results in sliding wear.
Figure 2.1a, shows the twin-disc rolling/sliding tribometer. The two circles on
top disc indicate that the top disc has radius of curvature whereas the bottom
disc has a flat surface. This will help for better alignment between the two discs
while conducting the experiments. This type of surface results in an elliptical
contact area. The two discs rotate with velocities V1 , and V2 , respectively, such
that V1 6= V2 . Such a system, from mechanics point of view, can be assumed
to be like the one shown on the right hand side of Figure 2.1b, in which the
bottom disc is fixed and the top disc rotates with the slip velocity. With this

12

2.2. Wear simulation strategy

assumption, the problem will be reduced from rolling/sliding contact to time


(pseudo) dependent sliding contact. Equation (2.9), computes wear depth for
nodes on the disc in a pin-on-disc tribometer. Therefore Equation (2.9) needs
to be modified, to compute wear in twin-disc rolling/sliding tribometer. The
modified wear model used in this work is explained in the following sub-section.

2.2.1.1

Generalized Archards wear model

Figure 2.2: Figure showing different positions assumed by reference node.

Consider a reference node, A, on the surface of the top disc, which comes in
contact with the bottom disc. As the disc rotates, the contact pressure on this
node increases from zero to maximum and then decreases gradually to zero (see
Figure 2.2). The node wears only when it experiences the pressure, i.e., when it
moves through the contact interface. Therefore, pressure acting on this point, is
integrated along the sliding distance which is the circumference of the disc. For
one rotation of the disc, the wear taking place on this node, on the top disc, can

2. Wear Simulation of Twin Disc Rolling/Sliding Tribometer

13

be written using the Equation (2.8) as,


Z

=2

hi+1 = hi + kD

pj R1i d,

(2.10)

=0

where is the circumferential coordinate, the subscript j corresponds to the


positions that the point occupies along its circumference while the disc rotates
through an angle of d , R1i is the radius of the top disc at the ith wear incre1 V2 )|
ment. For a given increment of time, ti , the top disc makes ti |(V
rotations.
2R1i
Where V1 and V2 are the velocities of the top disc and the bottom disc, respectively. Therefore the amount of wear depth, for a time increment of ti , is,

hi+1

ti | (V1 V2 ) |
= hi + kD
2R1i

=2

pj R1i d.

(2.11)

=0

This is the wear model used in this work. Equation (2.11) is called as generalized
Archards wear model. The wear depth is computed using Equation (2.11) for all
surface nodes.

In order to compute the integral in Equation (2.11), it is necessary that the nodes
on the surface of the FE model are located exactly along the circumference of
the various streamlines on the disc. A streamline is a linear path on the surface
of the top disc, traced by the surface nodes as the disc rotates. The width of the
disc can be assumed to be made of several streamlines. Therefore, the FE model
has been modeled in such a way that all the nodes, on the surface of the disc,
appear along various streamlines. Figure 2.3 shows the bottom view of the top
disc. This figure represents only a quarter of the top disc (read Section 3.1) as
this will save a lot of computation time and memory. We can see that various
nodes on the surface, which come in contact with the bottom disc, are located
along the streamlines. Figure 2.3 also shows the edge nodes and the corner nodes.
In this work, for the edge nodes and corner nodes the inward surface normal is
calculated in the user subroutine UMESHMOTION. More details of computation
of normals is given in Sub-section 2.3.2.1.

14

2.3. Wear simulation of twin-disc rolling/sliding tribometer.

Figure 2.3: Schematic representation, showing bottom view of the top disc.

2.3

Wear simulation of twin-disc rolling/sliding


tribometer.

The wear simulation of twin-disc tribometer using generalized Archards wear


model is accomplished using commercial FE code, ABAQUS 6.5, together with
user defined subroutine, UMESHMOTION (see Figure 2.4). Using adaptive
meshing algorithm, it is possible, in ABAQUS 6.5, to specify the mesh constraints through a user defined subroutine, UMESHMOTION. In this work, wear
depth computed using Equation (2.11), by UMESHMOTION is applied as mesh
constraint on all surface nodes. The user subroutine, UMESHMOTION coded
for simulating wear in twin-disc rolling/sliding tribometer is named as Wear-

2. Wear Simulation of Twin Disc Rolling/Sliding Tribometer

15

Contact Pressure,
Normals

A
B
A

User Subroutine,
UMESHMOTION

Q
U
S
Wear increment

Figure 2.4: Figure showing connection between ABAQUS and the user defined
subroutine, UMESHMOTION.
Simulator. In the following lines a detailed description of adaptive meshing and
working of Wear-Simulator is given

2.3.1

Adaptive meshing

In structural analysis, sometimes, as the deformation takes place, some of the elements are overly distorted and may not produce good results. So it is advisable
that those elements are re-meshed near those regions. This can be accomplished
by using adaptive meshing which re-locates the mesh, without creating new elements in the domain. This property of the adaptive meshing algorithms is used
for wear simulation to re-locate the mesh by an amount equal to the computed
wear depth, 4h. In ABAQUS 6.5, adaptive meshing combines the features of
pure Langrangian analysis and Eulerian analysis (ABAQUS, 2004). That is, the
structural analysis is carried out using pure Lagrangian approach and the relocation (also called as mesh smoothing) of the distorted elements is carried out
using Eulerian approach. Thus adaptive meshing maintains high quality mesh
throughout the analysis, even when large deformation or loss of material occurs,
by allowing the mesh to move independently of the material. Adaptive meshing consists of two fundamental tasks, (1) creating new mesh, through a process
called as sweeping, (2) remapping the solution variables from the old mesh to
the new mesh through a process called as advection.

16

2.3. Wear simulation of twin-disc rolling/sliding tribometer.

2.3.1.1

Mesh sweeping

Adaptive mesh smoothing is performed after the structural equilibrium equations


have converged. The mesh smoothing equations are solved explicitly by sweeping,
iteratively, over the adaptive mesh domain. During each mesh sweep, nodes in the
domain are re-located, based on the positions of the neighboring nodes obtained
during the previous mesh sweep, to reduce element distortion. The new position,
xj+1 , of a node is obtained as,
xj+1 = X + uj+1 = NN xN
j ,
where X is the original position of the node, uj+1 is the displacement, xN
j are the
neighboring nodal positions obtained during the previous mesh sweep, and NN
are the weight functions obtained from a least squares minimization procedure
that minimizes displacement in the original configuration (ABAQUS, 2004) .

2.3.1.2

Advection

During advection, the solution variables are re-mapped from the old mesh to the
new mesh by integrating the advection equations using the Lax-Wendroff method
(Leveque, 2002). The scheme is explicit, second order accurate and provides some
upwinding. Upwinding is a behavior of the solution in which it remains constant
along the characteristic curves. For a linear advection equation as shown below,
qt + aqx = 0,
the solution is constant along the lines dx
= a, and these lines are called as
dt
characteristic curves (for more details, refer chapter 3 and 4 of Leveque (2002)).
The Lax-Wendroff scheme for linear advection equation is,
qjn+1 = qjn

(at)2 n
at n
n
n
),
)+
(qj+1 qj1
(q 2qjn + qj1
2x
2x2 j+1

where the subscript n stands for the time increment and j stands for the node
number in the adaptive mesh domain, q is the solution variable, being advected
e.g., contact pressure, displacement etc., a is advection velocity, t and x are
the increments in time and space respectively. The advection of material quantities can lead into loss of equilibrium, for two main reasons. The first reason

2. Wear Simulation of Twin Disc Rolling/Sliding Tribometer

17

is, the error in the advection process itself. To minimize the errors in advection,
ABAQUS restricts the magnitude of the advection velocity by taking Courant
number ( for more details about Courant number and numerical stability please
refer Gustavsson (2004)) to be less than one. The second reason for loss of equilibrium is the changes in the representation of the underlying material quantities
by the changed mesh. This effect is more severe in case of wear simulation. Because, as the disc wears, the contact area increases and hence more number of
elements come in contact. This will lead to slightly different volume of integration while computing the internal forces. Therefore ABAQUS solves the last time
increment, to correct the equilibrium.

2.3.2

Wear-Simulator

It is to be noted that in wear simulation of twin-disc tribometer, the top disc,


which is defined as deformable body, is not rotated physically, but it is assumed
to be rotated for certain time increment in the user defined subroutine UMESHMOTION which we call the Wear-Simulator. During this time increment, it is
assumed that the configuration changes are negligible and have minor effect on
the contact solution. Bottom disc, which is defined as rigid body, is rotated by a
small angle, physically, to include the frictional effects as shown in Figure ??.
First, the three dimensional, deformable-rigid contact problem is solved in
ABAQUS. After the equilibrium equations have converged, ABAQUS invokes
the adaptive meshing algorithm. Adaptive meshing algorithm, further, invokes
the user defined subroutine UMESHMOTION. The working of Wear-Simulator
will be explained with the help of a flow-chart, shown in Figure 2.5.
The Wear-Simulator accesses the nodes one-by-one and stores them in a global
matrix in the order of their appearance on the bottom surface of the top disc.
Then, nodal coordinates and respective contact pressures are accessed. The pressure experienced by the node changes as it moves through the contact interface.
Therefore pressure is integrated over the sliding distance for each streamline. To
compute the number of rotations in Equation (2.11) the circumference of the
disc, for that increment, is calculated . Wear depth in this time increment is
calculated using Equation (2.11) for all surface nodes. The inward surface nor-

18

2.3. Wear simulation of twin-disc rolling/sliding tribometer.

Geometry
Contact Definition
Material Model
Load
Boundary Condition

Finite Element Simulation


(ABAQUS)

Wear-Simulator

Surface Nodes
Coordinates
Contact Pressure

Yes

Inc=1

Surface
Node Map

No
Integrate Pressure
Circumference
Local Wear Model
(Generalised Archards
Wear Model)
Wear Depth

Node on
Edge

Yes

Inward Surface
Normal

No
Wear in the Direction of
Inward Surface Normal

Sweep the Mesh


Advect

Adaptive Meshing

t tmax

Yes

No
END

Figure 2.5: Flowchart for wear simulation in a twin-disc tribometer.

2. Wear Simulation of Twin Disc Rolling/Sliding Tribometer

19

mals for all surface nodes are supplied by ABAQUS, whereas for the nodes which
come on edges and corners (see Figure 2.3), the inward surface normals have to
be computed. Computing the inward normals for edge nodes and corner nodes
is discussed in the Subsection 2.3.2.1. The wear depth for all surface nodes are
returned to the adaptive meshing algorithm. Wear depth, received by adaptive
meshing algorithm (see Sub-section 2.3.1) for all surface nodes, are applied in two
steps. First, the surface nodes are swept in the inward normal direction. Sweeping of the nodes is carried out purely as Eulerian analysis. Thus the geometry
is updated. Second, the material quantities are re-mapped to the new positions.
This is accomplished by advecting the material quantities from the old location to
the new location by solving advection equations using a second order numerical
method, called as Lax-Wendroff method as explained above. The advection of
material quantities can cause equilibrium loss, as explained in Subsection 2.3.1.2.
The equilibrium loss is corrected by solving the last time increment. Thus the
contact pressure is updated. The process from invoking the adaptive meshing
algorithm to advecting the material point information is called one wear increment. It is to be noted that, in this work, the complete contact problem is not
solved at the end of each wear increment. Instead, only the last time increment
is solved to correct the loss of equilibrium. This will save a lot of computation
time compared to wear simulation tool, developed in Hegadekatte et al. (2005).

2.3.2.1

Computation of inward surface normals

For nodes which are on the surface the inward surface, normals are supplied by
ABAQUS. But for nodes which are on the edges as well as those which are on
the corners (Figure 2.3) the normals have to be computed. In this section a brief
description of computing inward surface normals is given (for more details refer
Hegadekatte et al. (2005)). The Figure (2.6) shows the node number 1, which is
on one of the edges of the FE model.

20

2.3. Wear simulation of twin-disc rolling/sliding tribometer.

4
s1
1
2
y

n 1
1

n 1
2

r1

r2

1
n 2
1

n 2
2

Figure 2.6: Computation of inward surface normals for edge nodes.

The node number 2 and 3 are to its left and right respectively, and node number
4 is a subsurface node. All these nodes lie on y z plane (see Figure 2.6) and
that is why the x coordinate for these nodes can be neglected while calculating
the inward normal. The inward surface normal at node 1 can be calculated using
the following condition,
ri nki = 0,

(2.12)

where the subscript i stands for the element number and superscript, k, stands
for the normal number of these elements at node 1. Applying this condition we
get four normals at node 1, out of which two will be acting outwards and the
remaining two will be acting inwards. Identification of the two normals which are
acting inwards is accomplished by imposing the condition that the dot product
of the inward normals and the vector formed by the subsurface node and node
under consideration, is always greater than or equal to zero. i.e.,
nki si 0,

(2.13)

where si is the subsurface vector formed between node 1 and node 4. With the
help of this condition we get two normals which are acting inwards. The unit
inward normal from these inward normals is calculated as,
ni =

(n1i + n2i )
.
k n1i + n2i k

(2.14)

2. Wear Simulation of Twin Disc Rolling/Sliding Tribometer

21

For the nodes which come on the corner, the inward normal is approximated
by but the unit normal in the direction of subsurface node. For example, in
Figure 2.7, for nodes 1 and 2 the inward surface normal is in the direction of
the respective subsurface nodes. A vector from the corner node to its subsurface
node is computed and is divided by its magnitude to get the unit vector in that
direction. Wear is applied in the direction of this unit vector.

1
n1

2
n2

y
z

Figure 2.7: Figure showing direction of wear for corner nodes.

In this chapter we discussed about the formulation of contact problem using FEM
and wear model used for simulating wear in twin-disc rolling/sliding tribometer.
A detailed description of wear model used in this work, adaptive meshing and
working of Wear-Simulator was also given. In the next chapter, the FE model
used for wear simulation and the results obtained from Wear-Simulator are presented.

Chapter 3
Wear Simulation Results
In this chapter, the results obtained using the Wear-Simulator to simulate wear
in a twin-disc rolling/sliding tribometer will be presented. First, the FE model
used for this purpose and the parameters used for the wear simulation will be
presented. Later the results of wear simulation will be presented.

3.1

FE model of the twin-disc rolling/sliding tribometer

The geometric model used for wear simulation is shown in Figure 3.2. In this
work a three dimensional FE model has been used for two reasons:
1. The top disc has a curved surface and the bottom disc has a flat surface,
as shown in Figure 3.2. Therefore the contact area will be an ellipse. Axisymmetric FE models cannot be used to represent an elliptical contact.
2. In (Hegadekatte, 2006b), it was shown that, for sliding contact, the asymmetric effects resulting from friction can be considerable.
Only a quarter of the top disc has been modeled (the shaded region in Figure
3.2), since the disc is symmetric about z-axis. Very fine mesh has been generated
near the contact region and the mesh is coarser as we go away from the contact

24

3.1. FE model of the twin-disc rolling/sliding tribometer

Figure 3.1: Geometric model of the twin-disc rolling/sliding tribometer.

region (see Figure ?? (a) and (b)). This will not only reduce, substantially, the
computation time and memory required, but will also result in more data points
for calculating the integral in the Equation (2.11). The aspect ratio of elements
near the contact region is very close to one. The FE model has been built in such
a way that each surface node appears along a streamline (Figure 2.3). This will
help in integrating pressure with respect to the sliding distance.
First, the mesh is created in 2-D in x-y plane. This mesh is then swept about
the center of the disc, to obtain a 3-D mesh. The bottom disc has been defined
as a rigid surface (analytical rigid surface in ABABQUS 6.5). The wear results
are obtained from the deformable top disc. The bottom disc is rotated by an
angle of 0.01 radians to include the frictional effects. The parameters used for
simulation are listed in the Table 3.1. In the absence of wear coefficient data from
twin-disc rolling/sliding tribometer experiments, it is assumed to be the same as
identified from pin-on-disc experiments (Herz et al., 2004; J. Schneider & Herz,
2005). Therefore dimensional wear coefficient, kD , and the frictional coefficient,
, are taken from pin-on-disc experimental data.

3. Wear Simulation Results

25

Table 3.1: Wear parameters for the twin-disc rolling/sliding tribometer (see Figure 3.2).
Parameter
Normal Load
Velocity of Wheel 1
Velocity of Wheel 2
Youngs Modulus
Poissons ratio
Frictional Coefficient
Dimensional Wear Coefficient

Value
FN = 300, 600, 900 mN
V1 = 800 mm/s
V1 = 880 mm/s
E1 = E2 = 152 GP a
1 = 2 = 0.32
= 0.6
kD = 1 1010 mm3 /N mm

Figure 3.2: Three dimensional picture of FE model of the top disc with zoomed
view showing the fine mesh near the contact region.

26

3.2. Results

0.004
0.003
Elastic + Wear

0.002

h [mm]

Wear-Simulator

0.001
0
0

6e+06

1.2e+07

N [-]

1.8e+07

2.4e+07

Figure 3.3: Graph of wear depth versus number of rotations.

3.2

Results

The wear simulations were carried out until 3 m of wear was obtained on the
top disc. Figure 3.3 shows the graph of wear depth, h, versus the number of
rotations, N . As the disc surface wears, the contact area, Ar , increases and hence
the pressure, p, decreases. Therefore the slope of the h versus N continuously
decreases as seen in Figure 3.3. This is called as running in, in the literature.
Unlike for wear depth curve, the curve for the elastic displacement and wear depth
starts at a non-zero value equal to the elastic displacement. It can be see that
this curve approaches the wear depth curve with increasing number of rotations.
The curve for the elastic displacement plus wear depth, is shown in Figure 3.3.
Figure 3.4 (a) and (b) shows how the contact pressure changes, when plotted on
surface nodes along x and z directions respectively. We can see that the pressure
drops along x and z directions as the disc wears out. However, as the disc wears
out, because of the geometry of the contacting surfaces, the semi-major axis, a,
of the elliptical contact area increases and the semi-minor axis, b, of the contact
area decreases (see Figure 3.5 (a) and (b)). With the increase in wear on the disc,
the elliptical contact slowly approaches towards a line contact. But the contact
area, on the whole, increases continuously as shown in t Figure 3.6 (a). Figure
3.6 (b) shows the decrease in pressure as the wear progresses on the disc.

3. Wear Simulation Results

27

(a)
x [mm]
0.05

0.1

0.15

-0.05

0.2

-150

p [N/mm ]

p [N/mm ]

0
0

(b)
z [mm]

-300

After
0.0E+00 Rotations
After
1.5E+05 Rotations
After
4.7E+05 Rotations
After 20.5E+05 Rotations
After 203.8E+05 Rotations

-450
-600

-0.025

0.025

0.05

-150

-300

-450

After 0.0E+00 Rotations


After 1.5E+05 Rotations
After 4.7E+05 Rotations
After 20.5E+05 Rotations
After 203.8E+05 Rotations

-600

Figure 3.4: Graph showing decay in contact pressure as wear progresses: (a) for
nodes located along x-axis; (b) for nodes located along z-axis.
(b)
0.024

0.15

0.018

b [mm]

a [mm]

(a)
0.2

0.1

0.012
0.006

0.05
0
0

6e+06

0
0

1.2e+07 1.8e+07 2.4e+07

N [-]

6e+06

1.2e+07 1.8e+07 2.4e+07

N [-]

Figure 3.5: (a) Graph showing increase in length of semi-major axis of the contact
area; (b) Graph showing decrease in length of semi-minor axis of the contact area.
(a)

(b)
600

p [N/mm ]

0.0015

Ar [mm ]

0.002

0.001

0.0005
0
0

6e+06 1.2e+07 1.8e+07 2.4e+07

N [-]

450
300
150
0
0

6e+06 1.2e+07 1.8e+07 2.4e+07

N [-]

Figure 3.6: (a) Graph showing increase in contact area; (b) Graph showing decrease in contact pressure as the disc wears.

28

3.3. Wear simulation with different loads and slips

3.3

Wear simulation with different loads and slips

As mentioned earlier, the experimental results for twin-disc rolling/sliding tribometer were not available. Therefore, in this work, two sets of simulations were
carried out, i.e., by varying the applied load with a constant slip and by varying
the slip with a constant load. With the help of these simulations, the time taken
by the twin-disc tribometer in the laboratory, to obtain a wear of 3 m, has been
approximated.

3.3.1

Wear Simulations with different loads

0.003
0.002
FN = 0.3 N
FN = 0.6 N
FN = 0.9 N

h [mm]

0.004

0.001
0
0

6e+06

1.2e+07

N [-]

1.8e+07

2.4e+07

Figure 3.7: Graph showing wear depth versus number of rotations for different
applied loads (slip = 10%).

Figure 3.7 shows the graph of wear depth versus number of rotations. The simulations are carried out for three different applied loads with a slip of 10%. As
the load increases the number of rotations required for 3 m of wear on the disc
decreases. The number of rotations shown on the x-axis of the Figure 3.7 are
for the system shown in Figure 2.1b, in which the top disc rotates with the slip

3. Wear Simulation Results

29

velocity and the bottom disc is rotated through a small angle (Sub-section 2.2.1).
The duration for which the experiments with the twin-disc tribometer need to be
conducted with different loads, for obtaining 3 m of wear on the disc, is listed
in Table 3.2.
Table 3.2: Time required for 3 m of wear in twin-disc tribometer for different
loads (see Figure 3.7).
Load (N)
0.3
0.6
0.9

3.3.2

Time (hours)
174
91
61

Wear Simulations with different slips

0.003
0.002

h [mm]

0.004

Slip = 5 %
Slip = 10 %
Slip = 20 %

0.001
0
0

4e+06

8e+06

t [s]

1.2e+07 1.6e+07

Figure 3.8: Graph showing wear depth versus number of rotations for different
slips (load = 0.3N).

Figure 3.8 shows the graph of wear depth versus number of rotations. The simulations are carried out for three different slips with a load of 0.3N . The time

30

3.4. Comparison with GIWM

shown on the x-axis of the Figure 3.8, are for the system shown in Figure 2.1b,
in which the top disc rotates with the slip velocity and the bottom disc is rotated
through a small angle (Sub-section 2.2.1). The duration for which the experiments with the twin-disc tribometer need to be conducted for obtaining 3 m of
wear on the disc, is listed in Table 3.3.

Table 3.3: Time required for 3 m of wear in twin-disc tribometer for different
slips (see Figure 3.8).
Slip
5%
10%
20%

3.4

Time (hours)
182
174
172

Comparison with GIWM

0.0045

GIWM
UMESHMOTION

0.003

h [mm]

0.006

0.0015
0
0

6e+06

1.2e+07

N [-]

1.8e+07

2.4e+07

Figure 3.9: Graph showing the comparison of wear results between GIWM and
the Wear-Simulator.

3. Wear Simulation Results

31

The results obtained from the Wear-Simulator are compared with Global Incremental Wear Model (GIWM) (Hegadekatte, 2006b). In GIWM, while calculating
the contact area for a simple geometry as in the case of the twin disc tribometer,
the elastic displacements are taken into consideration by using the Oliver & Pharr
(1992) relation. The average contact pressure (global quantity) is calculated since
the applied load is known. Then, the global incremental wear is calculated and
integrated over sliding distance. Figure 3.9 shows the graph of wear depth obtained from both GIWM and the Wear-Simulator. It can be seen from the graph
that the wear calculated by Wear-Simulator is less than that of GIWM. The difference in results between the two, is 35.63%. To look further into this difference
the results obtained from Wear-Simulator are discussed by comparing them with
another simulation tool, Wear-Processor, in the next chapter.

Chapter 4
Discussion and conclusion

In this chapter wear results obtained from the Wear-Simulator are discussed in
comparison with GIWM and Wear-Processor Hegadekatte et al. (2005). To evaluate more closely, a test model with a coarse mesh, to save computational time,
was used, in the same way as the main model. Since the test model has coarse
mesh, in order to obtain reasonable number of nodes in contact, a higher load
was applied. Also, the dimensional wear coefficient used, was hundred times that
of the value used for FE model discussed in previous chapter (main model). The
remaining parameters like the velocities of the two discs, frictional coefficient,
Youngs modulus and Poissons ratio remain the same as those used for main
model. The wear parameters used for test model are listed in the Table 4.1.
Table 4.1: Wear parameters for test model.
Parameter
Normal Load
Velocity of Disc 1
Velocity of Disc 2
Youngs Modulus
Poissons ratio
Frictional Coefficient
Dimensional Wear Coefficient

Value
FN = 300 N
V1 = 800 mm/s
V1 = 880 mm/s
E1 = E2 = 152 GP a
= 0.32
= 0.6
kD = 1 108 mm3 /N mm

34

0.0300
0.0225

Wear-Processor

0.0150

Wear-Simulator

h [mm]

GIWM

0.0075
0.0000
0

1500

3000

N [-]

4500

6000

Figure 4.1: Graph showing wear depth obtained from three different tools:
GIWM, Wear-Processor and Wear-Simulator.
Figure 4.1 shows the linear wear obtained from three different tools, namely,
Wear-Simulator, Wear-Processor and GIWM. The working of Wear-Simulator
and GIWM has been discussed in chapter 2 (Sub-section 2.3.2) and chapter 3
(Section 3.4) respectively. And working of Wear-Processor is explained in brief
in Section 2.2. In Figure 4.1 we can see that the results from GIWM and WearProcessor match each other, favorably. The wear calculated by Wear-Simulator
is less than that of Wear-Processor. The difference in wear depth between WearProcessor and Wear-Simulator is 17.46 %. In Figure 4.2 variation of pressure for
the two simulation tools is shown.
One of the differences between the two simulation tools, namely the WearProcessor and the Wear-Simulator, is in calculating the contact pressures. In
the Wear-Processor, the contact pressures are calculated, for each surface node,
by using the stress tensors and normals. Whereas in the Wear-Simulator contact
pressures are accessed, directly, from the main code of ABAQUS. There are two
principal differences between the two simulation tools. First one, is in updating the pressure field at the end of every wear increment. As discussed earlier,
in the Wear-Processor the complete contact problem is solved, whereas, in the
Wear-Simulator, pressure field is updated by correcting the equilibrium loss. The
second one, is in the numerical integration method used in these two simulation

4. Discussion and conclusion

35

4000

Wear-Simulator
Wear-Processor

p [N/mm ]

4500

3500
3000
2500
2000
0

1500

3000

N [-]

4500

6000

Figure 4.2: Graph showing decrease of pressure obtained by two different simulation tools: the Wear-Processor and the Wear-Simulator.
tools. In Wear-Processor, the numerical integration method used for integrating
pressure, p, with respect to sliding distance along the streamlines, was Simpsons
1/3rd rule and in Wear-Simulator it was mid-point rectangular rule.
In the next section of this chapter, the results from two simulation tools: WearSimulator and Wear-Processor, are discussed with respect to numerical integration method used. In the later section, updating of the pressure field by ABAQUS,
is discussed. In the last section a conclusion is given.

4.1

Numerical integration method

In Wear-Processor, the numerical integration method used for integrating pressure, p, with respect to s, along each the streamlines, was Simpsons 1/3rd rule
and in Wear-Simulator it was mid-point rectangular rule. When the difference
between these two methods was studied, it was seen that for the first streamline,
before wear, there is a difference of 16.58 %, as shown in Table 4.2.

36

4.1. Numerical integration method


Table 4.2: Comparison of integration methods used.
Integration Method
Simpsons 1/3rd rule (Wear-Processor)
Mid-point rectangular rule (Wear-Simulator)

Value
1170 N/mm
980 N/mm

The wear depth calculated and applied by Wear-Processor is more than that
of Wear-Simulator as the integration of pressure lies in the numerator of the
Equation (2.11). This could be one of the reasons why the pressure (Figure 4.2),
as a function of number of rotations, drops faster for Wear-Processor than that
of the Wear-Simulator.

0.0300
0.0225

Wear-Processor

0.0150

Wear-Simulator (Simpsons rule)

h [mm]

GIWM

0.0075
0.0000
0

Wear-Simulator (Mid-point rectangular rule)

1500

3000

N [-]

4500

6000

Figure 4.3: Graph showing improved wear results on the test model from WearSimulator by using Simpsons 1/3rd rule in comparison to Wear-Processor and
GIWM.

4. Discussion and conclusion

37

Wear-Simulator
Wear-Processor

4000

p [N/mm ]

4500

3500

Using Simpsons rule for integration

3000
2500
2000
0

1500

3000

N [-]

4500

6000

Figure 4.4: Graph showing the pressure drop obtained from Wear-Simulator on
test model by using Simpsons 1/3rd rule in comparison to that obtained from
Wear-Processor.

In order to compare the results obtained from the Wear-Processor and the WearSimulator, Simpsons 1/3rd rule was implemented in Wear-Simulator as well. It
can be seen from Figure 4.3 that using Simpsons 1/3rd rule the wear depth obtained from Wear-Simulator are higher than that of wear depth obtained from
mid-point rectangular rule. Also, Figure 4.4 shows that the pressure drop obtained from Simpsons 1/3rd rule is faster than the pressure drop obtained from
mid-point rectangular rule. The difference in wear depth between the WearProcessor and the Wear-Simulator is 8.26 % and the pressure difference between
the two is 3.9 %. From Figure 4.3 and 4.4 we can see that by using a better
integration scheme the wear results have improved. However, for the FE model
considered in the previous chapter, there are sufficient number of data points to
integrate pressure over a streamline. Therefore, there is a negligible difference
between the two different integration methods, as seen in Figure 4.5.

38

4.2. Pressure field updated by ABAQUS

0.0015

GIWM
Wear-Simulator
Using Rectangular rule

0.001
Using Simpsons rule

h [mm]

0.002

0.0005
0
0

1.5e+06 3e+06 4.5e+06 6e+06

N [-]

Figure 4.5: Graph showing wear depth obtained on main model using two different integration methods.

4.2

Pressure field updated by ABAQUS

4000

p [N/mm ]

4500

3500
With pressure correction

3000
Without pressure correction

2500
2000
0

1500

3000

N [-]

4500

6000

Figure 4.6: Graph showing pressure updated by ABAQUS 6.5 using test model.

4. Discussion and conclusion

39

As discussed earlier ABAQUS does not solve the complete contact problem at
the end of a wear increment, but only the last increment is solved to correct
the equilibrium loss. Therefore, at the end of certain number of rotations the
pressure updated by ABAQUS were verified. Simulation on the test model using
Simpsons 1/3rd rule, was terminated after 2000 rotations, and the pressure field
was updated by solving the complete contact problem. Figure 4.6 shows the
updated contact pressure as a function of number of rotations. It can be seen
that, the pressure updated by ABAQUS does not agree, completely, with the
pressure obtained by solving the complete contact problem. The difference of
7.54 % is present between the two after 2000 rotations. However, this difference
is negligible compared to the saving in the computation time. Wear simulation
was continued with the Wear-Simulator using the corrected pressure field, for
another 1300 rotations. In this simulation, pressure decreased faster. This is
because, with the corrected pressure field, the calculated wear was more and
hence there was increase in contact area. Similarly, pressure was corrected after
3300 rotations and the wear simulation was continued with this new pressure field.
Figure 4.7 shows wear depth results obtained after the pressure is corrected. We
can see that there is a marginal increase in the wear obtained after pressure
correction. Similarly pressure field was corrected at different intervals using main
model. From Figure 4.8 and Figure 4.9, we can see that the corrected pressure
field has produced negligible amount of increase in wear depth.
0.04

With pressure correction

0.02
Without pressure correction

h [mm]

0.03

0.01
0.00
0

1500

3000

N [-]

4500

6000

Figure 4.7: Graph showing change in wear depth after pressure correction, using
test model

40

4.2. Pressure field updated by ABAQUS

500

p [N/mm ]

Wear-Simulator

400
Without Pressure Correction
With Pressure Correction

300
200
100
0

2.5e+05

5e+05

N [-]

7.5e+05

1e+06

Figure 4.8: Graph showing pressure updated by ABAQUS 6.5 using main model.

h [mm]

0.0005
GIWM

0.0004
0.0003
0.0002

With Pressure Correction

Without Pressure Correction

0.0001
0
0

2.5e+05 5e+05 7.5e+05 1e+06

N [-]

Figure 4.9: Graph showing change in wear depth after pressure correction, using
main model

4. Discussion and conclusion

4.3

41

Conclusion

On the basis of the results presented so far, the following conclusions are made:
1. It was shown that the user defined subroutine UMESHMOTION (WearSimulator) in ABAQUS can be coded to simulate rolling/sliding wear in a
twin-disc tribometer.
2. The Wear-Simulator is faster than that of the Wear-Processor. It was observed that, for the test model, the time taken by Wear-Simulator was 40
minutes, whereas the time taken by Wear-Processor for the same amount of
wear, on the same computer, was 600 minutes. However, the computation
time required by GIWM is negligible.
3. The wear results obtained from Wear-Simulator are compared with GIWM
and a difference of 35.63 % was observed.
4. A better numerical integration method, like Simpsons 1/3rd rule, was used
and it was seen that this does not produce significant improvement on the
wear results.
5. Pressure field updated by ABAQUS at the end of every wear increment
is not very accurate. A difference of 7.54 % was observed between the
pressure updated by ABAQUS by correcting only the equilibrium loss and
pressure obtained by solving the complete contact simulation. However,
this difference is difference is negligible when the saving in the computation
time is considered.
6. A difference of 8.26 % was observed, in the wear depth between the WearProcessor and the Wear-Simulator using Simpsons 1/3rd rule, which is negligible. Since Wear-Simulator is computationally less expensive compared
to the Wear-Processor, a best compromise between computation time and
accuracy has to be made in using the two FE based wear simulation tools.

Chapter 5
Summary
In this work, an attempt has been made to simulate wear in a twin-disc
rolling/sliding tribometer with defined slip. A wear simulation strategy in association with FEM (Hegadekatte et al., 2005) has been used. User defined
subroutine UMESHMOTION (Wear-Simulator) in ABAQUS has been applied,
as an alternative tool, to simulates dry sliding wear between the two discs. Wear
simulations were carried out for different applied loads and different slips and
the results obtained from these simulations were presented. Wear depth obtained
from Wear-Simulator were compared with GIWM and a difference of 35.63 % was
found. Results obtained from Wear-Processor, GIWM and Wear-Simulator, on a
test model, were discussed with respect to their numerical integration methods. It
was shown that, use of a better numerical integration scheme, like Simpsons 1/3rd
rule, improves the results on the test model. However, use of Simpsons 1/3rd rule
does not produce significant improvement on the main model as the main model
has sufficient number of data points for integrating pressure with respect to sliding distance for each streamline. The pressure updated by ABAQUS was verified
and a difference of 7.54 % was observed which is negligible. Unlike in WearProcessor, where the complete contact problem is solved at the end of every wear
increment, in this work the adopted method (UMESHMOTION/Wear-Simulator)
updates the contact pressure by correcting the equibrium loss by solving the last
time increment of the contact simulation. This saves a lot of computation time.
Therefore Wear-Simulator is computationally less expensive compared to WearProcessor. However, a difference of 8.26 % in the wear results from the two tools,
43

using test model with Simpsons 1/3rd rule, is present. A best compromise between computation time and accuracy has to be made, in using the two FE based
wear simulation tools. This work forms the basis of the future work, which aims
to simulate wear in micro-machines and thus predict useful life-span.

List of Figures
1.1

(a) Picture showing linear rack micro-motor, (b) Picture showing


gap formation on the gear, because of wear. Courtesy: Dannelle
M. Tanner (Sandia National Laboratories). . . . . . . . . . . . . .

1.2

Twin disc rolling/sliding tribometer. . . . . . . . . . . . . . . . . .

2.1

(a) Schematic view of the twin-disc tribometer; (b) Figure showing


the contact simulation geometric model for twin-disc tribometer. .

11

2.2

Figure showing different positions assumed by reference node. . .

12

2.3

Schematic representation, showing bottom view of the top disc. .

14

2.4

Figure showing connection between ABAQUS and the user defined


subroutine, UMESHMOTION. . . . . . . . . . . . . . . . . . . . .

15

2.5

Flowchart for wear simulation in a twin-disc tribometer.

. . . . .

18

2.6

Computation of inward surface normals for edge nodes. . . . . . .

20

2.7

Figure showing direction of wear for corner nodes. . . . . . . . . .

21

3.1

Geometric model of the twin-disc rolling/sliding tribometer. . . .

24

3.2

Three dimensional picture of FE model of the top disc with zoomed


view showing the fine mesh near the contact region. . . . . . . . .

25

3.3

Graph of wear depth versus number of rotations. . . . . . . . . . .

26

3.4

Graph showing decay in contact pressure as wear progresses: (a)


for nodes located along x-axis; (b) for nodes located along z-axis.

27

3.5

3.6

(a) Graph showing increase in length of semi-major axis of the


contact area; (b) Graph showing decrease in length of semi-minor
axis of the contact area. . . . . . . . . . . . . . . . . . . . . . . .

27

(a) Graph showing increase in contact area; (b) Graph showing


decrease in contact pressure as the disc wears. . . . . . . . . . . .

27

45

46

LIST OF FIGURES
3.7 Graph showing wear depth versus number of rotations for different
applied loads (slip = 10%). . . . . . . . . . . . . . . . . . . . . .
3.8 Graph showing wear depth versus number of rotations for different
slips (load = 0.3N). . . . . . . . . . . . . . . . . . . . . . . . . .

29

3.9 Graph showing the comparison of wear results between GIWM and
the Wear-Simulator. . . . . . . . . . . . . . . . . . . . . . . . . .

30

4.1 Graph showing wear depth obtained from three different tools:
GIWM, Wear-Processor and Wear-Simulator. . . . . . . . . . . .

34

4.2 Graph showing decrease of pressure obtained by two different simulation tools: the Wear-Processor and the Wear-Simulator. . . . .
4.3 Graph showing improved wear results on the test model from
Wear-Simulator by using Simpsons 1/3rd rule in comparison to
Wear-Processor and GIWM. . . . . . . . . . . . . . . . . . . . . .
4.4 Graph showing the pressure drop obtained from Wear-Simulator
on test model by using Simpsons 1/3rd rule in comparison to that
obtained from Wear-Processor. . . . . . . . . . . . . . . . . . . . .
4.5 Graph showing wear depth obtained on main model using two

28

35

36

37

different integration methods. . . . . . . . . . . . . . . . . . . . . 38


4.6 Graph showing pressure updated by ABAQUS 6.5 using test model. 38
4.7 Graph showing change in wear depth after pressure correction,
using test model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.8 Graph showing pressure updated by ABAQUS 6.5 using main model. 40
4.9 Graph showing change in wear depth after pressure correction,
using main model . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

List of Tables
3.1
3.2

Wear parameters for the twin-disc rolling/sliding tribometer (see


Figure 3.2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Time required for 3 m of wear in twin-disc tribometer for different

25

loads (see Figure 3.7). . . . . . . . . . . . . . . . . . . . . . . . .


Time required for 3 m of wear in twin-disc tribometer for different
slips (see Figure 3.8). . . . . . . . . . . . . . . . . . . . . . . . . .

30

4.1

Wear parameters for test model. . . . . . . . . . . . . . . . . . . .

33

4.2

Comparison of integration methods used. . . . . . . . . . . . . . .

36

3.3

47

29

48

LIST OF TABLES

Bibliography
ABAQUS (2004). V 6.5. Hibbit, Karlsson and Soresen Inc., Providence, RI,
USA.
Bathe, K. J. (1996). Finite element procedures In engineering Analysis. Prentice
Hall.
Deuflhard, P. (2004). Newton Methods for Non-linear Problems, Affine Invariance
and Adaptive Algorithms, volume 35. Springer,Berlin.
Gustavsson, K. (2004). Computational fluid dynamics. Lecture Notes.
Hegadekatte, V. (2006a).

Global incremental wear model for twin disc

rolling/sliding tribometer. Personal Communication.


Hegadekatte, V. (2006b). Modelling and simulation of dry sliding wear for micromachine applications. Doctoral Thesis.
Hegadekatte, V., Huber, N., & Kraft, O. (2005). Finite element based simulation
of dry sliding wear. Modelling Simul. Mater. Sci. Eng., 13, 5775.
Herz, J., Schneider, J., & Zum-Gahr, K. H. (2004). Tribologische charakterisierung von werkstoffen fr mikrotechnische anwendungen. In R. W. Schmitt
(Ed.), GFT Tribologie-Fachtagung 2004 Gttingen, Germany. on CD.
J. Schneider, K. H. Z.-G. & Herz, J. (2005). Tribological characterization of mold
inserts and materials. In D. Lhe & J. H. Hauelt (Eds.), Micro-Engineering
of Metals and Ceramics, part I and part II (pp. 579-603). Wiley-VCH Verlag
GmBH, Weinheim, Germany.
Leveque, R. (2002). Finite Volume Methods for Hyperbolic Problems. Cambridge
University Press.
49

50

BIBLIOGRAPHY

Oliver, W. C. & Pharr, G. M. (1992). An improved technique for determining


hardness and elastic modulus using load and displacement sensing indentation
experiments. J. Mat. Res., 7, 15641583.
Romig, A., Dugger, M. T., & McWhorter, P. J. (2003). Materials issues in
microelectromechanical devices: science, engineering, manufacturability and
reliability. Acta Materialia, 51, 58375866.
Spiegelberg, C. (2005). Friction and Wear in Rolling and Sliding Contacts. Doctoral thesis.
Williams, J. A. (2001). Friction and wear of rotating pivots in mems and other
small scale devices. Wear, 251, 965972.
Ypma, T. J. (1995). Historical development of the newton-raphson method.
SIAM, Review 37 (4), 531551.

Potrebbero piacerti anche