Está en la página 1de 17

Advanced friction modeling for sheet metal

forming

Authors
J.Hola, M.V. Cid Alfarob, M.B. de Rooijc, T. Meindersd
a

Materials innovation institute (M2i)


Tata Steel Research, Development & Technology
c
University of Twente, Faculty of Engineering Technology, Laboratory for Surface
Technology and Tribology
d
University of Twente, Faculty of Engineering Technology, chair of Forming Technology b

Please Note
Care has been taken to ensure that the information herein is accurate, but Tata Steel and its subsidiary companies do not accept
responsibility for errors or for information which is found to be misleading. Suggestions for or descriptions of the end use or
applications of products or methods of working are for information only and Tata Steel and its subsidiaries accept no liability in respect
thereof. Before using products supplied or manufactured by Tata Steel the customer should satisfy themselves of their suitability
All drawings, calculations and advisory services are provided subject to Tata Steel Standard Conditions available on request.

Advanced friction modeling for sheet metal forming


J.Hola,, M.V. Cid Alfarob , M.B. de Rooijc , T. Meindersd
a Materials innovation institute (M2i) - P.O. box 5008 - 2600 GA Delft - The Netherlands
Steel Research, Development & Technology - P.O. box 10000 - 1970 CA IJmuiden - The Netherlands
c University of Twente, Faculty of Engineering Technology, Laboratory for Surface Technology and Tribology - P.O. box 217 - 7500 AE Enschede - The Netherlands
d University of Twente, Faculty of Engineering Technology, chair of Forming Technology - P.O. box 217 - 7500 AE Enschede - The Netherlands
b Tata

Abstract
The Coulomb friction model is frequently used for sheet metal forming simulations. This model incorporates a constant coefficient
of friction and does not take the influence of important parameters such as contact pressure or deformation of the sheet material
into account. This article presents a more advanced friction model for large-scale forming simulations based on the surface changes
on the micro-scale. When two surfaces are in contact, the surface texture of a material changes due to the combination of normal
loading and stretching. Consequently, shear stresses between contacting surfaces, caused by the adhesion and ploughing effect
between contacting asperities, will change when the surface texture changes. A friction model has been developed which accounts
for the change of the surface texture on the micro-scale and its influence on the friction behavior on the macro-scale. This friction
model has been implemented in a finite element code and applied to a full-scale sheet metal forming simulation. Results showed a
realistic distribution of the coefficient of friction depending on the local process conditions.
Keywords:
friction modeling, friction mechanisms, asperity contact, flattening, real contact area, ploughing, adhesion

1. Introduction
The automotive industry makes extensive use of Finite Element (FE) software for formability analyses to reduce the cost
and lead time of new vehicle programs. In this respect, FE
analysis serves as a stepping stone to optimize manufacturing
processes. However, an accurate forming analysis of an automotive part can only be made if, among others, the material behavior and friction conditions are modeled accurately. For material models, significant improvements have been made over
recent decades. However, in the majority of simulations a simple Coulomb friction model is still used. This model does not
incorporate the influence of important parameters on the contact behavior, such as pressure, punch speed or deformation of
the sheet material. Consequently, even using the latest material models, it is still cumbersome to predict the draw-in and
springback of a blank during the forming process correctly.
To better understand contact and friction conditions during
lubricated sheet metal forming (SMF) processes, experimental and theoretical studies have been performed. At the microscopic level, friction is due to adhesion between contacting asperities [1, 2], the ploughing effect between asperities [1, 2]
and the appearance of hydrodynamic friction stresses [3, 4].
Ploughing effects between asperities and adhesion effects between boundary layers are the main factors causing friction in
the boundary lubrication regime. If the contact pressure is carried by the asperities and lubricant flow - as in the mixed lubri Corresponding

author. Tel.: +31 534894567; fax: +31 534893471.


Email address: j.hol@m2i.nl (J.Hol)

Preprint submitted to WEAR

cation regime - or fully carried by the lubricant - as in the hydrodynamic lubrication regime - hydrodynamic shear stresses will
contribute or even predominate. This article will focus on the
two friction mechanisms present in the boundary layer regime:
ploughing and adhesion.
Wilson [1] developed a model which treated the effect of adhesion and ploughing separately. A more advanced model was
developed by Challen & Oxley [2] which takes the combined
effect of ploughing and adhesion on the coefficient of friction
into account. Challen & Oxley performed a slip-line analysis on
the deformation of a soft flat material by a hard wedge-shaped
asperity and derived expressions for the coefficient of friction
and wear rates. Westeneng [5] extended the model of Challen
& Oxley to describe friction conditions between multiple tool
asperities and a flat workpiece material. Their model considers the flattened plateaus of the workpiece asperities as soft and
perfectly flat, and the surface texture of the tool as hard and
rough.
The amount of ploughing and adhesion depends on the real
area of contact. Hence, the coefficient of friction will change
if the real area of contact changes. The real area of contact
depends on the various flattening and roughening mechanisms
of the deforming asperities. The three dominating flattening
mechanisms during SMF processes are: (1) flattening due to
normal loading [6]; (2) flattening due to stretching [7, 8]; and
(3) flattening due to sliding [9]. Flattening increases the real
area of contact, resulting in a higher coefficient of friction.
Roughening of asperities, observed during stretching of the deformed material [10], tends to decrease the real area of contact
November 15, 2010

resulting in a lower coefficient of friction. The two mechanisms


outlined in this article are flattening due to normal loading and
flattening due to stretching. Future work is planned on modeling the roughening effect and the influence of sliding on the
flattening behavior.
A major research area within the field of friction modeling is
focused on developing models to predict the flattening behavior of asperities due to normal loading. Most of these models
are based on the pioneering work of Greenwood & Williamson
[6] which developed a stochastic model based on contact between a flat tool and rough workpiece surface. Their model
mathematically describes contact between two surfaces based
on the assumption that summits of the rough surface are spherical, that summits only deform elastically and that the surface
texture can be described by a distribution function. Over recent decades, modifications have been made to this model to
account for arbitrary shaped asperities, plastically deforming
asperities and the interaction between asperities. Zhao and
Chang [11] developed a model to describe interactions between
asperities on the micro-scale using the Saint-Venant principle
and Loves formula (elastic interaction between asperities is assumed). This model is integrated into the elastic-plastic contact
model of Zhao, Maietaa and Chang (ZMC-model) [12] which
includes a transition regime from elastic to fully plastic deformation of asperities. Jeng and Wang [13] extended the ZMCmodel for elliptical contact situations and Pugliese et al. [14]
for parabolic profile approximations. Pullen & Williamson [15]
developed an ideal plastic contact model based on the conservation of volume during plastic deformation and the assumption that displaced material reappears as a uniform rise in the
non-contacting surface. Conservation of energy is used to obtain relations between the contact pressure, separation and real
area of contact as a function of the surface height distribution.
The model of Pullen and Williamson inspired Westeneng [5]
to derive an ideal-plastic and nonlinear-plastic contact model
based on the conservation of volume and energy. Westeneng
modeled the asperities as bars which can represent arbitrarily
shaped asperities. The models include a persistence parameter,

work-hardening parameters and are able to describe the interaction between asperities.
A further increase of the real area of contact could occur if,
during normal loading, a bulk strain is applied to the material.
The effective hardness of the asperities can be largely reduced
if a bulk strain is present in the underlying material [7]. Wilson & Sheu [7] developed an analytical upperbound model to
describe the flattening behavior using wedge-shaped asperities
with a constant angle, Figure 1a. The length of the asperities
is much greater than the width of the asperities. Therefore, a
plane-strain state transverse to the asperities (x-direction) and a
plane-stress state in the direction of the asperities (y-direction)
is assumed since the stress in this direction might be neglected.
The semi-empirical relation of Wilson & Sheu provides a relation between the effective hardness, the real area of contact
and a non-dimensional strain rate. Sutcliffe [8] extended the
model of Wilson & Sheu to describe a plane-strain situation in
the direction of the asperities (strain in y-direction equals zero,
Figure 1b). A slip-line analysis is performed to describe the
flattening of transverse wedge-shaped asperities. Westeneng
[5] developed a strain model which describes the influence of
strain on a surface geometry using arbitrary shaped asperities.
His method is based on volume and energy conservation laws
and assumes that the crushed asperities cause a constant rise of
the non-contacting asperities. The model is applicable to both
plane-strain and plane-stress situations, depending on the definition of the non-dimensional strain rate [5].
In this article, a friction model is proposed which includes
various friction mechanisms: flattening due to normal loading,
flattening due to straining, ploughing and adhesion. Existing
models have been used to describe these mechanisms:
- To describe the flattening behavior of asperities due to normal loading the contact model of Westeneng [5] has been
used. His contact model include flattening parameters
which are not included in other loading models. Therefore, the model of Westeneng will likely have better predicting capabilities in describing the flattening behavior of

FN

FN

y
y , y = 0

y = 0

y
x
x

x = 0
(a)

(b)

Figure 1: Representation plane-stress model of Wilson and Sheu (a) and plane-strain model of Sutcliffe (b)
2

2. The friction model starts with defining the process variables


and material characteristics (step 1). Process variables are the
nominal contact pressure and strain in the material which are
calculated by the FE code. The contact force carried by the
asperities equals the total nominal contact force since hydrodynamic friction stresses will not be accounted for. Significant
material characteristics are the hardness of the asperities and
the surface properties of the tool and workpiece material. Once
the input parameters are known, the real area of contact is calculated based on the models accounting for flattening due to
normal loading (step 2) and flattening due to stretching (step
3). The amount of indentation of the harder tool asperities into
the softer workpiece asperities can be calculated if the real area
of contact and the contact pressure carried by the asperities are
known (step 4). After that, shear stresses due to ploughing and
adhesion effects between asperities (step 5) and the coefficient
of friction (step 6) are calculated. It is noted that in reality flattening due to normal loading and flattening due to stretching
will appear simultaneously during sheet metal forming, as well
as the combination between flattening (step 2 and 3) and sliding (step 4 and 5). Nevertheless, it has been assumed that the
various mechanisms act independently of each other in this research.
Friction models encompassing micro-mechanisms are generally regarded as too cumbersome to be used in large-scale
FE simulations. Therefore, translation techniques are necessary
to translate microscopic contact behavior to macroscopic contact behavior. Using stochastic methods, rough surfaces are described on the micro-scale by their statistical parameters (mean
radius of asperities, asperity density and the surface height distribution). Statistical parameters can be used if assumed that
the surface texture is isotropic and can be represented by 2dimensional correlated random noise. It is assumed that these
restrictions are true for the workpiece and tool material, which
makes the use of statistical parameters desirable in making the
translation from micro to macro contact modeling. Assuming
that the surface height distribution on the micro-scale represents
the surface texture on the macro-scale, it is possible to describe
contact problems that occur during large-scale FE analyses of
sheet metal forming processes [5].

1. Input step: Process variables and


material characteristics

2. Flattening due to normal loading


(Section 2.3.1)

3. Flattening due to stretching


(Section 2.3.2)

4. Indentation model (Section 2.4)

5. Calculation shear stresses


(Section 2.4)

6. Calculation coefficient of friction


(Section 2.4)

Figure 2: Solution procedure


asperities than other models.
- The flattening behavior of asperities due to straining has
been described by the strain model of Westeneng [5]. The
strain model of Westeneng includes flattening parameters which are not accounted for in other models. Especially the possibility to describe arbitrarily shaped asperities makes the strain model preferable to others.
- The influence of ploughing and adhesion on the coefficient
of friction has been described by the extended model of
Challen & Oxley [2, 5]. The ability to describe contact
problems between multiple asperities makes this extended
version favorable.

2.2. Characterization of rough surfaces


A discrete surface height distribution of the tool and workpiece material is obtained from surface profiles (Figure 3a).
However, a continuous function is desirable to eliminate the
need for integrating discrete functions during the solution procedure of the friction model. Various methods exist to describe
discrete signals by continuous functions. The Gauss distribution function can be used if it is assumed that the surface height
distribution is symmetric and approximates a normal distribution function. However, the initial surface height distribution
is usually asymmetric and will become even more asymmetric
if there is flattening of contacting- and rising of non-contacting
asperities. The asymmetric Weibull distribution function is a
more flexible criterion but can only approximate smooth surface height distributions. A more advanced method to describe
discrete signals can be achieved by using a Fourier series. A

An overview of the friction model is presented in this article


and the translation from micro to macro modeling is outlined.
The theoretical background of the models used to describe the
various friction mechanisms is described and the implementation in FE codes is discussed. The flattening models are validated by means of FE simulations on the micro-scale and the
applicability of the friction model in FE codes is proven by a
full-scale sheet metal forming simulation.
2. Theoretical background
2.1. Unified friction theory
A friction model, to be used in finite element codes, has been
developed to couple the various micro friction models, Figure
3

160

Real distribution

133

107
(z)(m1 )

z(m)

Surface profile

80

-2

53.3

-4

26.7

-6

0.8

2.4

1.6

3.2

Fourier fit half range

-4

-2.4

0.8

-0.8

x(mm)

2.4

z(m)
(b)

(a)

Figure 3: Surface profile (a) and corresponding surface height distribution (b)
Fourier series makes it possible to describe non-smooth asymmetric distribution functions from which the accuracy of the
evaluation depends on the number of expansions used.
The results discussed in this article are obtained by evaluating the surface height distribution functions (z) by a half range
sine Fourier function [16], given by:
(z) =

X
n=1

bn sin

 n 
z
L

rough surface (z), the uniform rise of the non-contacting surface U (based on volume conservation) and the separation between the tool surface and the mean plane of the asperities of
the rough surface d.
2.3.1. Flattening due to normal loading
Using the normalized surface height distribution (z), the
amount of flattening of the contacting asperities d and the rise
of the non-contacting asperities U can be calculated based on
energy and volume conservation. Contact between a flat hard
smooth surface and a soft rough surface is assumed without
sliding and bulk deformation. Only plastic deformation of asperities is assumed without work-hardening effects.
The amount of external energy must equal the internal energy in order to account for energy conservation. The amount
of external energy is described by the energy needed to indent
contacting asperities. The internal energy is described by the
energy absorbed by the indented asperities and the energy required to lift up the non-contacting asperities. The indentation variables used to describe Westenengs model are depicted
in Figure 5. A distinction is made between asperities in contact with the indenter, asperities which will come into contact
due to the rise of asperities and asperities which will not come
into contact with the indenter. The amount of indentation is
described by the variable z while the rise of asperities is described by the variable u. The number of asperities in contact
with the indenter is indicated by the counter N with corresponding indentation heights of zi (i = 1, 2, ..., N). The number of
asperities coming into contact with the indenter due to a rise of
non-contacting asperities is described by the counter N with
indentation heights of z j ( j = 1, 2, ..., N ). Hence, the total
number of asperities in contact with the indenter after applying
the normal load equals N + N . Asperities which will not come
into contact during the load step are indicated by the counter
N with corresponding rising heights of ul (l = 1, 2, ..., N ).

(1)

with:
2
bn =
L

ZL
0

f (z) sin

 n 
z
L

(2)

in which n represents the number of expansions, L the evaluation domain and f (z) the discrete form of the surface height
distribution. In Figure 3b, the discrete surface height distribution from the workpiece material (Figure 3a) is evaluated by a
Fourier function using 15 expansions.
2.3. Flattening mechanisms
Two flattening mechanisms have been implemented in the
friction model to calculate the real area of contact of the workpiece: flattening due to normal loading and flattening due to
stretching. The models of Westeneng are used for this purpose
[5]. Westeneng assumed the tool as rigid and perfectly flat,
which indents into a soft and rough workpiece material. This
assumption is valid since the difference in hardness and length
scales between the tool and workpiece material is significant in
the case of sheet metal forming processes. Westeneng modeled
the asperities of the rough surface by bars which can represent
arbitrarily shaped asperities, Figure 4. Westeneng introduced
3 stochastic variables as presented in Figure 4: The normalized surface height distribution function of the asperities of the
4

Rise non-contacting
Indented asperities

Surface U

Tool surface
d

Mean plane

(z)

Workpiece
asperities
U

Figure 4: A rough soft surface indented by a smooth rigid surface


The total amount of asperities M equals N + N + N .
The amount of external energy depends on the total number
of asperities in contact with the indenter (N + N ). Normally,
the non-contacting asperities would rise with a distance ul .
But due to the presence of the indenter some of the asperities
are restricted to rise with a distance of u j . A certain amount of
external energy is required to prevent a rise of z j = ul u j .
The energy required to indent contacting asperities is given by:
Wext =

N
X

F Ni zi +

i=1

N
X

F N j z j =

j=1

N+N
X

F Nk zk

The amount of internal energy is described by the summation


of the energy absorbed by the indented asperities Wintab and the
energy required to rise the non-contacting asperities Wintri over
all asperities, Equation 5.
Wint = Wintab + Wintri

The asperities are described by bars having the same area


A. The maximum pressure an asperity can carry before deformation occur equals the hardness H of the material since ideal
plasticity is assumed. Therefore, the absorbed energy Wintab
over N + N indented asperities can be described as:

(3)

k=1

Equation 3 can be rewritten as:

dk
Wext = 1 F N z

with

1 =

Wintab =
N+N
X

N+N
X

HAzk

(6)

k=1

F Nk zk

k=1

Wintab can also be written as Equation 7 using Ar = (N + N ) A


representing the real area of contact and 2 representing a shape
factor.
N+N
X
zk
k=1
dk
Wintab = 2 HAr z
with
2 =
(7)
dk
(N + N )z

(4)

dk F N
z

dk the maximum
In which F N represents the total force and z
indentation height. 1 is called the energy factor since it is influenced by the amount of external energy required to indent
the rough surface.
F Ni

(5)

Wintri is described by the sum of energy required to rise N


asperities which comes in contact with the indenter after application of the normal load and N asperities which do not come
into contact with the indenter, Equation 8:

N
N
X

X
ul HA
(8)
Wintri = u j HA +

N
zi

j=1

z j

Equation 8 include a persistence parameter which describes


the amount of energy required to lift up the non-contacting asperities. A value of = 0 means that no energy is needed to
rise the asperities, a value of = 1 implies that a maximum
amount of energy is needed to rise the asperities. Equation 8
can be simplified by using volume conservation:

u j
V

l=1

ul

Wintri = H

Figure 5: Zoom-in on indented and rising asperities

N
X
i=1

zi A

(9)

which can also be written as:

dk
Wintri = 3 H (Ar N A) z

with

3 =

N
X

constant rise of asperities U are unknowns in Equation 16. An


equation can be obtained from volume conservation, Equation
17. Using stochastic parameters, the equation for volume consistency can be written as Equation 18:

zi

i=1

dk
N z

(10)
N
X

3 is called the shape factor since it is influenced by the shape


of the surface.
Balancing the total internal energy (Equation 5) to the external energy (Equation 4) gives:
2
3
FN
= Ar + (Ar N A)
H
1
1

Anom

3 = 2 = 1

(12)

F N (z) (z) dz

N+N = M

U (1 ) =

u (z) (z) dz + Anom

zk d
zk d + U

(15)

In which the surface height distribution of the rough surface


(z), to be used for the translation from micro to macro friction
modeling, has been introduced. Substituting Equation 12 into
Equation 11 gives:

Zd
Z

Pnom

(z) dz = 1 + (z) dz (16)


= +

H
dU

(18)

z1

(z d) (z) dz

(19)

2.3.2. Flattening due to stretching


Besides an ideal-plastic contact model for normal loading,
Westeneng derived an analytical contact model to describe the
influence of strain on deforming, arbitrary shaped, asperities.
The effective hardness of asperities can be largely reduced if a
bulk strain is present in the underlying material. As a result,
more indentation of contacting asperities will occur. The outcome of the ideal-plastic load model, described in Section 2.3.1,
will be used as an input. The subscript S has been used to indicate the variables which are strain dependent. For example,
US corresponds to the rise of non-contacting asperities and dS
to the indentation of contacting asperities due to straining.
Analogous to the loading model, the stretching model considers contact between a flat hard surface and a soft rough surface. Only plastic material behavior is assumed without workhardening effects. The derivation of the stretching model starts
on single asperity scale, where the change of the fraction of
the real contact area S as a function of the nominal strain is
derived.
d S
d can be written as a function of the indentation speed of
S
the indenting asperity d
dtS and the strain rate in the underlying
bulk material dtds :

(14)

for zk > d
for d U zk < d

(d z) (z) dz

The amount of flattening d of contacting asperities and the


rise of non-contacting asperities U due to normal loading can
be calculated by solving Equation 16 and Equation 19 simultaneously.

(13)

(z) dz = M

Zd

dU

and:
(

Zz1

where z1 represents the initial height of an asperity which just


comes into contact with the indenter after applying the normal load, (z1 + u(z1 ) = d). Taken a constant rise U of the noncontacting asperities into account, equation 18 becomes:

dU

zk =

(z d) (z) dz =

Anom

dU

represent the ratio of the real to the nominal area of contact,


can be regarded as a energy factor and as a shape factor. , as
well as and are variables which depend on the statistical parameters U (the constant rise of asperities) and d (the separation
between the tool surface and the mean plane of the asperities of
the rough surface). In addition, is a function of the normal
forces acting on the asperities F N (z). It should be noted that a
constant rise of asperities has been assumed in the derivation
of , and , which corresponds to the experimental results
of Pullen & Williamson [15]. Statistical parameters have been
introduced by using the following stochastic variables, with M
the total amount of asperities:
Z

(17)

u j A

j=1

FN = M

N
X

| {z } | {z }

(11)

ul A +

l=1

| {z }

The parameters 1 , 2 and 3 can be expressed as a function of


each other. Appendix A shows the derivation of the expressions
given in Equation 12.
2 =

zi A =

i=1

N
X

Equation 16 gives the relation between the nominal contact


pressure Pnom (defined as F N /Anom ), the separation d and the
constant rise of the non contacting asperities U. Another equation is required to compute Pnom since the separation d and the

d S dS dtS
d S
=
d
dS dtS d
6

(20)

S represents the indentation distance of an asperity. The


increase of the indentation distance dS can be written as:
dS = |dUS | + |ddS | = d (US dS )

To calculate the change of S , the value of US and dS needs


to be solved simultaneously while is incrementally increased.
Based on volume conservation (Equation 31) and the definition
of the fraction of real contact area (Equation 28) US and dS can
be obtained.

(21)

Hence, the first term in Equation 20 can be written as:


d S
d S
=
dS
d (US dS )

(22)

US (1 S ) =

dtS
1
=
d

2.4. Calculation of shear stress


The model of Challen & Oxley [2, 17] takes the combining
effect of ploughing and adhesion between a wedge-shaped asperity and a flat surface into account. Westeneng [5] extended
the model of Challen & Oxley to describe friction conditions
between a flat workpiece material and multiple tool asperities,
Figure 6b. He assumed that the flattened peaks of the asperities
are soft and perfectly flat and the surface of the tool material is
rough and rigid. As mentioned earlier, the difference in hardness between the tool and workpiece material and the difference
in length scales between the two surfaces is significant in the
case of a sheet metal forming process. Therefore, it is valid to
make a subdivision in two length scales using a rigid tool and a
soft workpiece. The macro-scale model of Challen & Oxley
has been implemented in the friction model to describe friction
conditions between the tool and workpiece material.

(23)

(24)

A non-dimensional strain rate can be defined as [7]:


E=

l
va + vb

(25)

with l representing the mean half spacing between asperities:


l=

2 QS

(26)

2.4.1. Shear stresses single asperity contact


The model of Challen & Oxley has been used to calculate
the coefficient of friction between a hard asperity and a soft flat
surface. The model takes the combining effect of ploughing and
adhesion between contacting surfaces into account. They performed a slip-line field analyses to describe contact conditions
of wedged-shaped asperities. The attack angle of the asperity
can be described by a constant angle [2] or a varying angle if
a spherical shaped asperity is assumed [17], Figure 6c. In case
of spherical shaped asperities, the attack angle is small if the
amount of indentation is small and grows if the amount of
indentation increases. This conforms more to reality then using
a constant attack angle. The amount of indentation and the
attack angle for a spherical-shaped asperity is defined as:

The definition for l, with Q representing the asperity density,


is approximately true for surfaces with no particular roughness
distribution [5].
Equation 24 can now be written as a function of the nondimensional strain rate:
d S
l
d S
=
d
E d (US dS )

(27)

It is assumed that the fraction of the real contact area for one
asperity S equals the total fraction of contact area S . Therefore, the stochastic form of the real contact area (Equation 28)
can be used to solve the differential equation in Equation 27:
S =

(z) dz

(28)

dS US

dS
d
=
d (US dS ) d (US dS )

(z) dz

dS US

(29)

= s

(32)

= = arctan p
2
(2t )

(33)

in which represents the distance between the mean plane of


tool asperities and the flat plateaus of the workpiece and t the
mean radius of tool asperities. The attack angle is chosen half
of the maximum attack angle which can be regarded as the
mean attack angle.

= (dS US )
Substituting Equation 29 into Equation 27 yields:
l
d S
= (dS US )
d
E

(31)

The implementation of Equations 30, 31 and 28 into FE


codes is shown in Section 3.2.

Substituting Equations 22 and 23 into Equation 20 gives:


d S
d S
va + vb
=
d
d (US dS )

(z dS ) (z) dz

dS US

The second term, which represents the velocity of the indenting asperity, is determined by the downward velocity of the indenting asperity va and the upward velocity of the rising asperities vb , Equation 23. The third term represents the strain in the
bulk material of the asperities and can be written as:
dS
= va + vb
dtS

(30)
7

Tool surface

Mean plane

F Nasp
t

t (s)
r
Workpiece surface

(a)

(b)

Boundary layer
(c)

Figure 6: Indentation tool asperities


2.4.2. Shear stress in a multi asperity contact
Westeneng described the translation from friction forces occurring at single asperity contacts to the total friction force at
multiple asperities by:

Challen & Oxley [17] deduced slip-line fields assuming a


plane-strain deformation state in the direction of the asperity
and ideal-plastic material behavior. The following set of equations has been found for the coefficient of friction asp for a
ploughing, wedge-shaped, asperity:
asp =

A sin + cos (arccos fC )


A cos + sin (arccos fC )

Fw = t S Anom

(34)

sin
A = 1 + + arccos fC 2 2 arcsin p

2
1 fC

(35)

(40)

In which the stochastic variable described in Equation 13 has


been used. t represents the asperity density of the tool surface,
S the ratio of the real to the nominal area of contact of the
workpiece, Anom the nominal contact area, t the normalized
surface height distribution function of the tool surface and Fwasp
the friction force occurring at a single asperity. The bounds of
the integral are described by smax , the maximum height of the
tool asperities, and , the separation between the workpiece surface and the mean plane of the tool asperities (Figure 6b). The
friction force Fw can be calculated once the amount of separation is known. can be obtained based on force equilibrium
for the normal load by solving Equation 41:

Fwasp = asp F Nasp

(36)

fC is called the friction factor defined as /k, with describing


the shear stress in the boundary layer and k the shear strength
of the softer material. The coefficient of friction depends on the
attack angle , and thus depends on the normal load F Nasp .
The friction force acting on one asperity is described by:

Zsmax
t (s) ds Pnom Anom
0 = HS Anom

The normal load one asperity can carry depends on the


amount of indentation . A relation between the radius of the
contact length r (Figure 6c) and the indentation can be defined as:
q
(37)
r = 2t 2

(41)

The integral in Equation 41 represents the fraction of the nominal contact area of the tool penetrating into the workpiece material.
If the shear stresses are known from Equation 40 the coefficient of friction can finally be obtained by:

The normal load becomes:


(38)

Assuming that only half of the contacting area carries the load
during ploughing and that only small indentations takes place
( << t ), F Nasp can be written as:
F Nasp = t H

Fwasp t (s) ds

with:



F Nasp = AH = 2t 2 H

Zsmax

Fw
FN

(42)

3. Implementation
The friction model described in Section 2.1 has been implemented into the in-house implicit FE code Dieka, developed at
the University of Twente. The friction model is called for every
node in contact during a FE simulation. If a node is in contact,
the nominal contact pressure and strain in the bulk material is

(39)

From which the friction force Fwasp acting on one asperity can
be calculated.
8

order to get values for the amount of flattening of contacting


asperities dS and the rise of the non-contacting asperities US ,
as presented in Figure 8.
The friction model can be used to solve plane-strain or
plane-stress situations, depending on the definition of the nondimensional strain rate E. Wilson & Sheu [7] developed a definition for E based on the flattening behavior of wedge-shaped
asperities using a constant angle. They assumed a plane-stress
state and a nominal strain in the longitudinal direction of the
asperities (y-direction) and a plane-strain state transverse to the
direction of the asperities (x-direction), Figure 1a. Based on
an upperbound analysis, Wilson & Sheu provided a relation between the effective hardness Hef f , the real area of contact and
the non-dimensional strain rate E. They also proposed a semiempirical relation for E as a function of H and S , fitted on the
results obtained by the upperbound-method [7]:

Pnom , H, and (z)

Pnom

0 = 1 + (z) dz

H
0=

(z d + U) (z) dz U

dU

, d and U

Figure 7: Calculation scheme load step

2
f2 (S )
Hef f
E=
f1 (S )

called from the source code. The discrete surface height distribution is described by a continuous function using the half
range Fourier serie (Section 2.2). Then, the fraction of real
contact area due to normal loading and stretching, shear stresses
due to ploughing and adhesion and the coefficient of friction are
being calculated by the equations described in Section 2. This
section describes the implementation of these equations into the
FE code in more detail.

E=

(43)

z =

=
zd
zd+U

for z > d
for d U z < d

Pnom
S k

(45)

1
0.184 + 1.21 exp (1.47)

(46)

with:

with:
(

Hef f =

in which f1 (S ) and f2 (S ) are fitting parameters. The parameter



k represents the shear strength of the surface material
H/3 3 .
Sutcliffe [8] extended the model of Wilson & Sheu to describe plane-strain situations. In their model a plane-strain state
in the longitudinal direction of the asperities (y-direction) and
a nominal strain transverse to the orientation of the asperities
(x-direction) is assumed, see Figure 1b. A slip-line analysis
has been performed to describe the flattening of wedge-shaped
asperities. Sutcliffe provided a relation between the effective
hardness Hef f , real area of contact S and non-dimensional
strain rate E. The relation between the non-dimensional strain
rate and the fan angle , based on results obtained by the slipline analysis, has been presented in their article. The fan angle
represents a characteristic slip-line angle which is bounded by
0 /2 due to geometrical conditions. Based on these results, Westeneng [5] proposed a semi empirical relation for the
non-dimensional strain rate as a function of the fan angle:

3.1. Implementation of flattening due to normal loading


The amount of flattening d of contacting asperities and the
rise of the non-contacting asperities U due to normal loading
can be calculated by solving Equation 16 and Equation 19 simultaneously. The solution scheme is schematically shown in
Figure 7.
An expression for F N (z) must be substituted into the expression for (see Appendix A, Equation A.3) in order to solve the
system of Equations presented in Figure 7. It is assumed that
the deforming asperities can be represented by plastically deforming spheres since an expression for F N (z) does not exist for
plastically deforming bars. Assuming that only small indentations occur (z << w ), the following expression for F N (z) can
be substituted (see also Equation 38):
F N (z) = 2w Hz

with

(44)

Hef f
(1 S )
4

(47)

The calculation scheme presented in Figure 8 can be solved


by implementing one of the two definitions of the nondimensional strain rate (Equation 45 or 46). The differential
equation can be solved by applying the Euler method while the
root of the inner two equations can be found by the second order
Newton Raphson method.

and w representing the mean radius of asperities. Using these


expressions the set of equations presented in Figure 7 can be
solved by using the second order Newton Raphson scheme.
3.2. Implementation of flattening due to normal loading and
stretching
Equation 30 has to be solved to calculate the change of the
fraction of real contact area S as a function of the nominal
strain . Equation 28 and 31 must be solved simultaneously in

3.3. Implementation calculation scheme for the coefficient of


friction
The amount of indentation of the harder tool asperities into
the softer workpiece asperities (Equation 41) can be calculated
9

dS
l
= (dS US )
d
E

U, d and from
load model

Pnom and in

S = S + dS

0=
Z

0=

(z) dz S

dS US

(z dS US ) (z) dz US

dS US

if + d < in

dS and US

Figure 8: Calculation scheme strain step, Pnom and in are coming in from the FE code
if the real area of contact is known from the two flattening
mechanisms. Shear stresses due to ploughing and adhesion effects between asperities and the coefficient of friction can be
calculated if the amount of indentation is known, as presented
in Figure 9.
An expression for the shear factor fC is necessary (Equation
34) in order to solve the system of Equations presented in Figure 9. fC is defined as /k with describing the friction force
in the boundary layer and k the shear strength of the softer material. The model of Timsit & Pelow [5, 18] has been imple-

mented in the friction model to describe the shear factor fC :


(p) = 3.94p0.81

During ploughing, the contact pressure p equals the effective hardness Hef f of the softer material since ideal plasticity
is assumed. The shear strength
k is related to the hardness H

by the relation k = H/3 3. Substituting Equation 48 and the


definition of k into the expression for fC gives:
fC =

Zsmax
0 = Hef f S Anom
t (s) ds Pnom Anom

The flattening models proposed by Westeneng are used to


determine the real area of contact between the tool and workpiece material. FE simulations on the micro-scale have been
performed in order to validate these models. Two sets of simulations have been performed for this purpose. In the first analysis, a two-dimensional rough surface of 4mm long was indented
by a perfectly flat and rigid tool, as shown in Figure 10a. The
second analysis was focused on indenting a rough surface by a
normal load including a bulk strain in the underlying material,
as shown in Figure 10b. The roughness distribution used in the
FE simulation equals the distribution measured for DC04 lowcarbon steel. The surface was modeled by 4 node 2D planestrain elements. The yield surface was described by the Von

Fwasp t (s) ds

(49)

4. Validation of flattening mechanisms

Fw = t S Anom

= 20.47Hef0.19
f
k

Timsit & Pelow performed experiments to obtain an empirical relation between the normal stress and the shear strength of
a stearic acid film deposited on aluminum. The relation is applicable for contact pressures in between 70MPa and 740MPa,
which are likely to occur on a micro scale during deep drawing
processes.

t (s), S , Pnom and Hef f

Zsmax

(48)

Fw
FN

Figure 9: Calculation scheme friction step


10

Tool

Tool

Workpiece

Workpiece

(VM stresses)

(VM stresses)

(a)

(b)

Figure 10: Schematic view analysis 1 (a) and analysis 2 (b)


Mises yield criterion using the Nadai hardening relation to describe work-hardening effects. Material parameters used for
the FE simulation are listed in Appendix B. Contact between
the tool and the rough surface was described by the penalty
method using a penalty stiffness of 1 N/mm. The surface height
distribution used for the analytical model corresponds to the
roughness distribution of the FE simulation. A fixed hardness
of 450 MPa (3y ) was used in the analytical model since a yield
strength of 150MPa was used for the FE simulation.
The amount of indentation of the rough surface - as well as
the development of the real area of contact - has been tracked
during the simulation and compared with the analytical solution. The analytical solution and the FE solution for elastic
ideal-plastic and elastic non-linear plastic material behavior is
presented in Figure 11 for analysis 1. The material behavior of
the elastic ideal-plastic FE simulation is comparable to the material behavior assumed in the analytical model (ideal-plastic).
The material behavior of the elastic nonlinear-plastic FE simulation corresponds more to the actual material behavior than
that of the elastic ideal-plastic FE simulation.
The analytical model uses a persistence parameter to
describe the amount of energy required to lift up the noncontacting asperities (see Equation 8). A value of = 0 means
that no energy is needed to rise the asperities, whereas a value
of = 1 implies that a maximum amount of energy is needed to
rise the asperities. Since the exact value of this parameter is not
known, different calculations have been performed in order to
obtain a precise value for this parameter. A higher value for the
persistence parameter results a lower amount of indentation
(Figure 11a) and a smaller value of the real contact area (Figure 11b). Both the indentation and the development of the real
contact area calculated by the analytical solution using a value
of = 1 correspond very well to the elastic ideal-plastic FE
solution. The analytical model deviates from the more realistic
elastic nonlinear-plastic FE simulation, since work-hardening
effects are not accounted for. The flattening of the asperities

will be lower due to work-hardening effects, which in turn result in a lower amount of indentation and real area of contact
(Figure 11).
Combined normal loading and stretching the underlying bulk
material decreases the effective hardness [7]. A lower hardness
results in an increase of the real area of contact. Both the analytical and the FE results of analysis 2, where a rough surface has
been indented by a nominal load and a bulk strain has been applied to the underlying material, are presented in Figure 12. For
the FE solution, only the development of the real area of contact is shown, since the deformation of the asperities is difficult
to separate from the deformation of the underlying bulk material. It can be concluded from Figure 12b that work-hardening
effects have a large influence on the flattening behavior of the
asperities. A difference of 20% in the real area of contact is
obtained at the end of the simulation between the results of the
elastic ideal-plastic and the elastic nonlinear-plastic simulation.
The non-dimensional strain rate E, used in the analytical
model, can be described by the definition given by Wilson &
Sheu (Equation 45) or the definition of Sutcliffe (Equation 46).
The results of the FE simulation presented in Figure 12b are
based on a plane-strain deformation mode in the longitudinal
direction of the asperities. Therefore, the definition proposed
by Sutcliffe has been used. The density of workpiece asperities (in mm2 ) is an unknown parameter in the analytical strain
model. Various methodologies exist to extrude the asperity density from the surface profile. Results obtained by these methodologies are highly dependent on the chosen method and the
resolution of the used roughness measurement device. Future
work is planned to determine this parameter using the most suitable method. Until then, the asperity density will be taken as an
unknown parameter. Calculations have been performed using
realistic values for the asperity density for DC04 to show the
importance of this parameter, see Figure 12. From Figure 12, it
can be concluded that the asperity density of the workpiece has
a significant influence on the development of the real area of
11

2.6

0.5
FEM: elastic ideal-plastic
= 0.5

Indentation (m)

2.2

= 1.0

Fraction of real contact area (-)

= 0.0

1.7
1.3
0.87
FEM: elastic ideal-plastic
FEM: elastic nonlinear-plastic

0.43

0.42

FEM: elastic nonlinear-plastic


Analytical solution

0.33

= 0.0

0.25

= 0.5
= 1.0

0.17
0.083

Analytical solution
0

20

40

80

60

100

20

Nominal contact pressure (MPa)

40

80

60

100

Nominal contact pressure (MPa)

(a)

(b)

Figure 11: Results analysis 1: Amount of indentation rough surface (a) and development of real area of contact (b)

FEM: elastic ideal-plastic

Fraction of real contact area (-)

5000 asp/mm

2.5
Indentation workpiece (m)

0.6

3000 asp/mm2

9000 asp/mm2

2
1.5
1

0.5

FEM: elastic nonlinear-plastic


Analytical solution

3000 asp/mm2

0.4

5000 asp/mm2
9000 asp/mm2

0.3
0.2
0.1

0.5
Analytical
0

0.02

0.04

0.06

0.08

0.1

0.02

0.04

0.06

0.08

0.1

Strain (-)

Strain (-)
(a)

(b)

Figure 12: Results analysis 2: Amount of indentation rough surface (a) and development of real area of contact (b)
contact. The amount of indentation of the workpiece asperities
will be lower if a higher value of the asperity density is used,
Figure 12a. Consequently, a lower amount of indentation results in a lower amount of real area of contact, Figure 12b. The
trend of the graphs corresponds well to the flattening behavior
obtained by the FE simulations. Using an asperity density of
5000 asp/mm2 it is possible to describe the results of the elastic
ideal-plastic FE solution (which has comparable material characteristics) precisely.

5. Application
The cross-die product is a test piece designed at Renault
which approximates process conditions of complex automotive
parts. The cross-die product is used to test the numerical performance of the developed friction model in a large-scale FE
simulation. In this article, the focus is on the numerical performance and feasibility of the friction model that has been developed. To validate this model, an experimental test procedure
needs to be developed and executed.
Due to symmetry of the cross-die product only a quarter of
the workpiece was modeled. The workpiece was meshed by
9000 triangular discrete Kirchhoff shell elements using 3 integration points in plane and 5 integration points in thickness
12

direction. The yield surface was described by the Vegter model


[19] using the Bergstrom-Van Liempt hardening relation [20]
to describe hardening behavior. Material parameters were used
from DC04 low carbon steel, a typical forming steel used for
SMF processes. A list of material parameters is listed in Appendix C. Contact between the tools and the workpiece was
described by a penalty method using a penalty stiffness of 200
N/mm. The coefficient of friction used in the contact algorithm
was calculated on the basis of the friction model presented in
this article. Roughness parameters are given in Table 1, and the
normalized surface height distribution functions that were used
for the tool and workpiece material are presented in Figure 13.
The simulation was performed by prescribing the displacement
of the punch until a total displacement of 50 mm was reached.
The punch speed was set to 5 cm/sec and the force applied to
the blankholder was 50 kN.
Two simulations have been performed in order to quantify
the individual contributions of the two flattening mechanisms.
The first simulation only accounted for the influence of normal
loading on the coefficient of friction, Figure 14. The second
simulation uses both flattening models to determine the coefficient of friction, Figure 15. Results shown in both Figure 14
and 15 are from the punch side of the sheet. The gray areas
represent the non-contacting areas.
If only flattening due to static loading is assumed (Figure
14), rather low values for the ratio of the real to the nominal
area of contact are obtained. This results in friction coefficients
that vary between 0.13 and 0.145. Higher values are obtained
in high-pressure regions: the contact area of the punch radius
(region A) and the thickened area of the blankholder region
(region B). Lower values occur in low-pressure regions: the
blankholder region and the top area of the punch. Results look
reasonable, but it should be noted that only one of the two flattening mechanisms was taken into account during the simulation. If the second flattening mechanism is taken into account
(flattening due to stretching), higher values for the real area of
contact are obtained (Figure 15). The higher contact ratios result in higher values of the coefficient of friction, i.e. between
0.13 and 0.19. It can be observed from Figure 15b that higher
values of the coefficient of friction occur at regions where high
strains occur (region C, D and E). Region C is purely stretched,

region D is compressed which causes thickening of the material and region E is stretched over the die radius. On the other
hand, low values of the coefficient of friction can be observed
in low-strain regions. Overall it can be concluded that the distribution of the coefficients of friction lies within the range of
expectation. The increase in calculation time is also promising. An increase of 60% for the first simulation and 200% for
the second simulation was obtained compared to the calculation
time required to perform a Coulomb based FE simulation of the
cross-die product.
During the implementation of the friction model into FE software some important assumptions had to be made:
- Full recovery of asperities is assumed during unloading
of the workpiece rough surface. The amount of recovery
of indented asperities is described by the amount of elastic spring-back in reality. Due to elastic spring-back, the
amount of recovery will be smaller than in case of full recovery. This will result a higher real area of contact at
lower loads. Consequently, the coefficient of friction will
be smaller due to a smaller amount of indentation of tool
asperities into the softer workpiece asperities. A realistic
unloading model is required to describe this effect.
- A definition of the non-dimensional strain rate is required
to calculate the amount of flattening due to bulk straining.
Various definitions exist to describe this variable, but most
of them are based on a plane-strain or a plane-stress assumption. The plane-strain definition of Sutcliffe (Section
3.2), taken the equivalent plastic strain as a strain measure,
has been used for the application discussed in this section.
However, strains and stresses occur in different directions
during sheet metal forming in reality. This gives rise to the
question if Sutcliffes model is still applicable and how the
strains should be accounted for.
- The model of Timsit & Pelow [5, 18] has been implemented in the friction model to describe the shear factor
fC (Section 3.3). Timsit & Pelow performed experiments
to obtain an empirical relation between the normal stress
and the shear strength of a stearic acid film deposited on
aluminum. The applicability of Timsit & Pelows model
0.3

Unit

Hardness workpiece (H)


Persistence parameter ()
Density workp. asp. (work )
Density tool. asp. (tool )
Radius tool. asp. (tool )
Nr. of Fourier expansions
Non-dim. strain rate (E)

1400
1
5.0 103
2.0 103
2.0 102
10
Sutcliffe

MPa

0.25

mm
mm2
mm

0.2

0.7

0.15

0.53

0.1

0.35

0.05

0.18

0
-4

Table 1: Roughness parameters

0.88

-2.4

-0.8

0.8

2.4

z(m)

Figure 13: Tool and workpiece distribution


13

(m1 ) tool

Value

(m1 ) workpiece

Roughness parameter

1.1
workpiece
tool

0.145

5.2%
A

0.0%

0.13

(b)

(a)

Figure 14: Development ratio of real to apparent area of contact (a) and coefficient of friction (b) for normal loading only (gray
represents the non-contacting area)
0.19

29.4%
C

0.13

0.0%
D
(b)

(a)

Figure 15: Development ratio of real to apparent area of contact (a) and coefficient of friction (b) for normal loading only (gray
represents the non-contacting area)
must be checked when using other metal-lubricant combinations.

influence of ploughing and adhesion effects between contacting


asperities on the coefficient of friction. A statistical approach is
adapted to translate the microscopic models to a macroscopic
level.
The friction model has been validated by means of FE simulations at a micro-scale. An excellent comparison between the
analytical and the FE simulation is obtained in case of indenting a rough surface by a normal load. It was also found that
work-hardening effects do not play a significant role in the case
of pure normal loading. If a nominal strain is applied to the bulk
material, the effect of work-hardening becomes much more significant. A difference of 20% in real contact area, compared

In an upcoming article, the validity of the friction model will


be shown - based on experimental research and large scale FE
simulations.
6. Conclusion
A friction model that can be used in large-scale FE simulations is developed. The friction model includes two flattening
mechanisms to determine the real area of contact at a microscopic level. The real area of contact is used to determine the
14

to the elastic ideal-plastic results, was found after applying a


nominal strain of 10%. In the case of a plane-strain mode in
the longitudinal direction of the asperities, the analytical model
is able to accurately describe the FE results based on a elastic ideal-plastic material model. However, it was not possible
to accurately describe the influence of work-hardening effects
due to the large difference between the elastic ideal-plastic FE
results and the elastic nonlinear-plastic FE results. It is concluded that work-hardening effects should not be neglected in
the analytical strain model. These effects must be included in
future models.
The friction model has been implemented in a FE code and
applied to a full-scale sheet metal forming simulation. Results
of the simulations have shown reasonable values for the coefficient of friction in the case of normal loading only, namely
between 0.13 and 0.145. If flattening due to stretching is also
incorporated, more realistic values are achieved (between 0.13
and 0.19).

13, 14 and 15 , 1 can be written in stochastic form:


Zd
1 =

F N (z) (z + U d) (z) dz +

F N (z) (z d) (z) dz

dU

Zd

(z + U d) (z) dz +

(z d) (z) dz

dU

2
Z

(A.3)

F N (z) (z) dz

dU

or:
2 =

(A.4)

with:
Ar
=
=
Anom

7. Acknowledgments

(z) dz

(A.5)

dU

This research was carried out under the project number


MC1.07289 in the framework of the Research Program of the
Materials innovation institute M2i (www.m2i.nl).

Derivation relation between 2 and 3


dk (Equation A.1) into EquaSubstituting the expression for z
tion 10 gives:

Appendix A.

3 = 2 (N + N )

This appendix presents the derivation of Equation 12 and the


expressions , and .

N
X

zi

i=1

N
X

(A.6)
zk

k=1

Derivation relation between 1 and 2

and in stochastic form, using N = M

(z)dz:

dk can be written as:


From Equation 7 z
dk =
z

N
X

zk

k=1

3 =

(A.1)

(N + N )2

1 =

dU

(z) dz
Z

(z d) (z) dz

Zd

F Nk zk

k=1

FN

(z) dz

dk into Equation 4, the definiSubstituting the expression of z


tion of 1 can be written as:
N+N
X

N+N
X

(N + N ) 2

(A.2)

dU

zk

(z + U d) (z) dz +

(A.7)
(z d) (z) dz

or:

k=1

3 = 2

Introducing the stochastic variables as presented in Equation


15

(A.8)

Appendix B.

[2] J. Challen, P. Oxley, An explanation of the different regimes of friction


and wear using asperity deformation models, Wear 53 (1979) 229243.
[3] N. Patir, H. Cheng, An average flow model for determining effects of
three-dimensional roughness on partial hydrodynamic lubrication, Journal of Lubrication Technology 100 (1978) 1217.
[4] W. Wilson, D. Chang, Low speed mixed lubrication of bulk metal forming
processes, Journal of Tribology 118 (1996) 8389.
[5] J. Westeneng, Modelling of contact and friction in deep drawing processes, Ph.D. thesis, University of Twente, 2001.
[6] J. Greenwood, J. Williamson, Contact of nominally flat surfaces, Proceedings of the Royal Society of London. Series A, Mathematical and
Physical sciences 295 (1966) 300319.
[7] W. Wilson, S. Sheu, Real area of contact and boundary friction in metal
forming, International Journal of Mechanical Science 30 (1988) 475489.
[8] M. Sutcliffe, Surface asperity deformation in metal forming processes,
International Journal of Mechanical Science 30 (1988) 847868.
[9] S. Lo, T. Yang, A new mechanism of asperity flattening in sliding contact
- The role of tool elastic microwedge, Journal of Tribology 125 (2003)
713719.
[10] H. Shih, W. Wilson, P. Saha, Modeling the influence of plastic strain on
boundary friction in sheet metal forming, Proc. NAMRC XXIV (1996)
173178.
[11] Y. Zhao, L. Chang, A model of asperity interactions in elastic-plastic
contact of rough surfaces, Journal of Tribology 123 (2001) 857864.
[12] D. M. Y. Zhao, L. Chang, An asperity microcontact model incorporating
the transition from elastic deformation to fully plastic flow, Journal of
Tribology 122 (2000) 8693.
[13] Y. Jeng, P. Wang, An elliptical microcontact model considering elastic,
elastoplastic and plastic deformation, Journal of Tribology 125 (2003)
232240.
[14] G. Pugliese, S. Tavares, E. Ciulli, L. Ferreira, Rough contacts between
actual engineering surfaces: Part II. Contact mechanics, Wear 264 (2008)
11161128.
[15] J. Pullen, J. Williamson, On the plastic contact of rough surfaces, Proceedings of the Royal Society of London. Series A, Mathematical and
Physical sciences 327 (1972) 159173.
[16] M. Greenberg, Advanced Engineering Mathematics - 2nd edition, Prentice Hall, 1998.
[17] J. Challen, P. Oxley, Slip-line fields for explaining the mechanics of polishing and related processes, International Journal of Mechanical Siences
26 (1984) 403418.
[18] R. S. Timsit, C. Pelow, Shear strength and tribological properties of
stearic acid films (part 1) on glass and aluminium-coated glass, Journal
of Tribology 114 (1992) 150158.
[19] H. Vegter, A. van den Boogaard, A plane stress yield function for
anisotropic sheet material by interpolation of biaxial stress states, International Journal of Plasticity 22 (2006) 557580.
[20] A. van den Boogaard, J. Huetink, Simulation of aluminium sheet forming
at elevated temperatures, Computer Methods in Applied Mechanics and
Engineering 195 (2006) 66916709.

Material parameters used for the FE simulations described in


Chapter 4
Material parameter

Value

Unit

Elastic modulus (E)


Poissons ratio ()
Initial strain (0 )
C-parameter (C)
n-parameter (n)

210
0.3
0.00243
500
0.2

GPa

MPa

Table B.2: Von Mises and Nadai parameters


Appendix C.
Material parameters used for the FE simulations described in
Chapter 5
Material parameter

Value

Unit

Elastic modulus (E)


Poissons ratio ()

210
0.3

GPa

Table C.3: Elastic material parameters


Material parameter

45

90

R-value
Uniaxial factor ( fun )
Plane-strain factor ( f ps )
Plane-strain ratio ()
Pure shear factor ( f sh )
Equi-biaxial factor ( fbi )
Equi-biaxial ratio (bi )

2.1537
1.0
1.317
0.5
0.5442
1.169
0.8427

1.3154
1.0531
1.3146
0.5
0.5846

2.1942
0.9979
1.3077
0.5
0.5295

Table C.4: Vegter parameters for DC04


Material parameter

Value

Unit

Initial static stress ( f 0 )


Stress increment parameter (dm )
Linear hardening parameter ()
Remobilization parameter ()
Hardening exponent (n)
Initial strain (0 )
Max. dynamic stress (0 )
Temperature (t)
Dynamic stress power (p)

114.8
250.34
0.25
8.1014
0.75
0.005
600
300
2.2

MPa
MPa

MPa
K

Table C.5: Bergstrom-Van Liempt hardening parameters for


DC04
References
[1] W. Wilson, Friction models for metal forming in the boundary lubrication
regime, Americal Society of Mechanical Engineers 10 (1988) 1323.

16

También podría gustarte