You are on page 1of 20

Engineering Computations

A solution of transient rolling contact with velocity dependent friction by the


explicit finite element method
Xin Zhao Zili Li
Article information:
To cite this document:
Xin Zhao Zili Li , (2016),"A solution of transient rolling contact with velocity dependent friction by the
explicit finite element method", Engineering Computations, Vol. 33 Iss 4 pp. 1033 - 1050
Permanent link to this document:
http://dx.doi.org/10.1108/EC-09-2014-0180
Downloaded on: 07 June 2016, At: 01:30 (PT)
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

References: this document contains references to 37 other documents.


To copy this document: permissions@emeraldinsight.com
The fulltext of this document has been downloaded 35 times since 2016*
Users who downloaded this article also downloaded:
(2016),"Using fuzzy logic control approach and model reduction for solving frictional contact
problems", Engineering Computations, Vol. 33 Iss 4 pp. 1006-1032 http://dx.doi.org/10.1108/
EC-11-2014-0226
(2016),"Turbulence models assessment for separated flows in a rectangular asymmetric three-
dimensional diffuser", Engineering Computations, Vol. 33 Iss 4 pp. 978-994 http://dx.doi.org/10.1108/
EC-05-2015-0112
(2016),"Using genetic algorithms to improve support vector regression in the analysis of
atomic spectra of lubricant oils", Engineering Computations, Vol. 33 Iss 4 pp. 995-1005 http://
dx.doi.org/10.1108/EC-03-2015-0062

Access to this document was granted through an Emerald subscription provided by emerald-
srm:235889 []
For Authors
If you would like to write for this, or any other Emerald publication, then please use our Emerald
for Authors service information about how to choose which publication to write for and submission
guidelines are available for all. Please visit www.emeraldinsight.com/authors for more information.
About Emerald www.emeraldinsight.com
Emerald is a global publisher linking research and practice to the benefit of society. The company
manages a portfolio of more than 290 journals and over 2,350 books and book series volumes, as
well as providing an extensive range of online products and additional customer resources and
services.
Emerald is both COUNTER 4 and TRANSFER compliant. The organization is a partner of the
Committee on Publication Ethics (COPE) and also works with Portico and the LOCKSS initiative for
digital archive preservation.
*Related content and download information correct at time of download.
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)
The current issue and full text archive of this journal is available on Emerald Insight at:
www.emeraldinsight.com/0264-4401.htm

A solution of transient rolling Explicit finite


element
contact with velocity dependent method
friction by the explicit finite
element method 1033
Received 22 September 2014
Xin Zhao Revised 26 July 2015
Department of Road and Railway Engineering, 12 October 2015
Accepted 1 December 2015
Delft University of Technology, Delft, The Netherlands and
Southwest Jiaotong University, Chengdu, People’s Republic of China, and
Zili Li
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

Department of Road and Railway Engineering,


Delft University of Technology, Delft, The Netherlands

Abstract
Purpose – The purpose of this paper is to develop a numerical approach to solve the transient rolling
contact problem with the consideration of velocity dependent friction.
Design/methodology/approach – A three dimensional (3D) transient FE model is developed in
elasticity by the explicit finite element method. Contact solutions with a velocity dependent friction law
are compared in detail to those with the Coulomb’s friction law (i.e. a constant coefficient of friction).
Findings – The FE solutions confirm the negligible influence of the dependence on the normal contact.
Hence, analysis is focussed on the tangential solutions under different friction exploitation levels. In the
trailing part of the contact patch where micro-slip occurs, very high-frequency oscillations are excited
in the tangential plane by the velocity dependent friction. This is similar to the non-uniform sliding or
tangential oscillations observed in sliding contact. Consequently, the micro-slip distribution varies
greatly with time. However, the surface shear stress distribution is quite stable at different instants,
even though it significantly changes with the employed friction model.
Originality/value – This paper proposes an approach to solve the transient rolling contact problem
with the consideration of velocity dependent friction. Such a problem was usually solved in the
literature by the simplified contact algorithms, with which detailed contact solutions could not be
obtained, or with the assumption of steady rolling.
Keywords Finite element method, Micro-slip, Rolling contact, Tangential contact solution,
Velocity dependent friction
Paper type Research paper

1. Introduction
Rolling contact can be found in rolling bearings, between wheels and rails, gears and
cam-followers (Sadeghi et al., 2009). Partial slip, or stick-slip, occurs when the contact is
not well lubricated. Constant kinematic coefficient of friction (COF) is often employed
for contact problems where friction is involved, e.g. in the widely referenced program
CONTACT of Kalker for frictional rolling (Kalker, 1990) and its extended version for
quasi-quarter spaces (Li and Kalker, 1998), and also in many other numerical or
analytical treatments of the tangential problem of rolling (Zhang et al., 2009; Lance and Engineering Computations:
International Journal for Computer-
Sadeghi, 1993; Xu and Jiang, 2002; Goryacheva et al., 2001). Furthermore, the difference Aided Engineering and Software
between the static and the kinematic COFs is often ignored. Vol. 33 No. 4, 2016
pp. 1033-1050
Friction in the nature, however, is often more complicated than Coulomb’s, and © Emerald Group Publishing Limited
0264-4401
frequently an accurate account of friction is necessary for a better and fundamental DOI 10.1108/EC-09-2014-0180
EC understanding of friction-related problems. Among the factors that affect COF, the
33,4 local relative velocity, i.e. micro-slip, of the contacting surfaces is often of particular
importance for rolling contact. It has been well known that a constant COF is not
sufficient to simulate the traction behavior of rail vehicles. For that, Polach (2005) has
included the velocity dependence of COF in a contact model developed on the basis
of a contact shear stiffness coefficient, while Giménez et al. (2005), Xie et al. (2006),
1034 Zhang et al. (2002) and Arias-Cuevas et al. (2010a, b) modified the FASTSIM of Kalker
to take it into account. For problems such as heat generation due to metal-to-metal
contact related to dynamic behavior of grease lubrication (Lugt, 2009), friction
management, control of wheel-rail traction, noise, and rolling contact fatigue
(Arias-Cuevas et al., 2010a, b; Dollevoet et al., 2010), velocity dependence of the
friction may also be important.
The approaches in Polach (2005), Giménez et al. (2005), Xie et al. (2006), Zhang
et al. (2002) and Arias-Cuevas et al. (2010a, b) are based on simplified contact
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

algorithms. They are fast and are suitable for calculation of resultant forces required by
multi-body dynamics, but cannot obtain detailed contact solutions such as
distributions of pressure, surface shear stress, stick-slip distinction, and so on.
Recently, Vollebregt and Schuttelaars (2012) extended Kalker’s full theory (i.e. program
CONTACT) to solve the 2-D rolling contact with velocity dependent friction, in which
the half-space assumption was employed and the inertia effects were ignored (i.e.
a quasi-static approach). Thus, 3-D solutions to transient rolling contact with the
consideration of velocity dependent friction still lack for realistic contact geometry.
This work intends to fill such a gap with a 3-D transient FE model that was initially
developed in Li et al. (2008) and validated for both normal and tangential contact
solutions in statics (Zhao and Li, 2011). The assumptions of half-space, linear elasticity
and steady state rolling, as are the basis of Kalker’s method, can all be dropped. In Zhao
and Li (2015) the FE model was also employed to examine the influence of plastic
deformation on detailed contact solutions. Necessity of a 3-D model for detailed
solutions of frictional rolling contact has been well recognized in the scientific
community, e.g. in researches into the rolling contact fatigue of bearings, gears,
cam-follower (Sadeghi et al., 2009). For the engineering purpose, Popescu et al. (2006)
developed a semi-analytical and a full analytical 3-D elasto-plastic contact models to
treat the normal contact problem of rolling bearings.
Instead of providing a fast creep force calculation method, this work aims at detailed
analyses of the relationships between friction laws, contact stresses, and micro-slip in
the contact patch and a better insight into the stick-slip phenomenon of rolling contact.
Different friction exploitation or creepage is considered. Here “stick” has the same
meaning as adhesion used in Kalker (1990) and Zhao and Li (2011), i.e. the opposite
state of slip. The approach is presented with wheel-rail contact in this work, but it is
also applicable to other types of rolling contact when the interface is not well lubricated
and metal-to-metal contact occurs (Lugt, 2009).

2. Modeling approach
In the 3-D transient FE model developed with ANSYS/Ls-dyna, the contact bodies are
meshed with eight-node solid elements, as shown in Figure 1, and the reduced Gaussian
integration is employed. With its application to wheel-rail rolling contact in this work,
the contact surface of the wheel is cylindrical, the radius of the rail profile at contact is
300 mm, and the contact surfaces are assumed to be smooth. This means that a
concentrated Hertzian contact occurs in the middle of the rail top against the wheel
(a) Mc
Sprung mass Explicit finite
element
K C
method


Wheel
z Solution zone
o x 1035
A B
Rail head H1

L1 L2

(b)
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

Wheel

Solution zone


Figure 1.
The 3D transient FE
Rail model of rolling
contact
Notes: (a) A schematic diagram; (b) the mesh

tread, through which the complexity in explaining the conformal contact phenomenon
is avoided and emphasis is placed on the influence of the velocity dependent friction.
The total element number is 219,048, and the minimum element size of 0.33 mm is
applied in the solution zone (an irregular mesh is employed to balance the model size
and the accuracy). A 3-D right-handed Cartesian coordinate system (Oxyz) is defined, of
which the origin O is located at the center of the solution zone (position B) where a
dense mesh is applied. The z axis is in the vertical direction.
In each simulation, after finding its equilibrium under gravity at the initial position
A with an implicit FE solver, the wheel is set to roll toward the solution zone with a
specified speed and a traction force, for which an explicit FE solver is employed.
The displacement field obtained from the implicit solver, the movement of the wheel in
rotation and forward translation are applied as the initial conditions of the explicit
simulation. The rolling distance AB is designed for the system to damp out the
influence of the initial conditions and approximately reach a steady state before
entering the solution zone, i.e. a dynamic relaxation is applied. During developing the
FE model the length of AB is determined by trial. The traction force is generated by
EC applying the corresponding driving torque to the wheel axle. A penalty method-based
33,4 surface-to-surface contact algorithm (Benson and Hallquist, 1990) is introduced to solve
the contact problem in the presence of friction and tangential load.
Since the explicit time integration is employed, a very small time step (2.6 × 10−8 s
for the model of this work) is required to meet the Courant et al. (1928) stability
condition, i.e. to ensure that a sound wave may not cross the smallest element during
1036 one time step. This ensures that the high-frequency dynamic effect is captured.
Boundary conditions are as follows: the rail head is supported continuously on a rigid
foundation; the ends of the rail heads are constrained along x-axis to take into account
its “infinite” length and the lateral ends of the wheel axle are constrained along y-axis to
include its symmetry. A linear-elastic material is used as listed in Table I to simulate
the shakedown state of the wheel and rail, i.e. after the running-in period of a new
system. Plastic deformation occurred during the shakedown process and the resulted
residual stresses (Popescu et al., 2006; Warhadpande et al., 2012) are not considered in
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

this work. Hereinafter, unless otherwise specified, contact solutions at instant t when
the center of the contact patch is at O are used for analyses. With a single 3.73 GHz
Pentium 4 processor, it takes about 27 h to run a typical simulation in which the wheel
rolls over a distance of 0.22 m under the default rolling speed of 140 km/h.
The FE model is further developed in the present work by considering the velocity
dependence of friction. At each node in contact, the local COF ( f ) is determined by the
following equation:

f ¼ f D þ ðf S f D Þeksn (1)

where fs is the static COF, fD is the kinematic COF, k is an exponential decay coefficient,
and sn is the relative velocity between particles in contact, i.e. the micro-slip at the node.
This is a basic assumption employed in this work, and the micro-slip is a function of the
creepages/spin and the rolling speed of the wheel over the rail. During the simulation

Parameters Values

Lumped sprung mass, Mc 13.4 t


Wheel diameter, ϕ 0.92 m
Wheel mass, Mw 900 kg
Rolling speed, v 140 km/h
Stiffness of primary suspension, K 1,150 KN/m
Damping of primary suspension, C 2,500 Ns/m
L1 150 mm
L2 30 mm
H1 36 mm
Wheel and rail material
Young’s modulus, E 210 GPa
Poisson’s ratio, υ 0.3
Density, ρ 7,800 kg/m3
Damping constant, β 1.0 × 10−4 s
Table I. Friction model
Values of Static COF, fs 0.5
parameters involved Kinematic COF, fD 0.32
in this work Decay coefficient, k 6
the local COF is updated at every time step. Note that a (slave) node is seldom in exact Explicit finite
contact with another (master) node in the simulation. Usually, it “contacts” with a element
(master) segment that is composed of 4 surface nodes. At each time step, the contact
segments are searched out for every slave node. The contact points, which are the
method
projection of the slave nodes on the corresponding contact segments, are also
determined. Results at a contact point are functions of those at the four nodes of the
contact segment. Here, sn is calculated from the velocities of the slave node and 1037
the corresponding contact point. For more information about the contact algorithm, the
readers are referred to Benson and Hallquist (1990).
Traction coefficient (μ) is distinguished from COF in this work by its definition of the
following equation:
m ¼ F L =F N p f r (2)
where FN and FL are the normal and longitudinal contact forces, respectively, and fr is
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

the overall COF of the contact pair that limits the traction force transmitted in the
contact. In each simulation, a constant driving torque corresponding to a specified FL is
applied to the wheel axis. No lateral friction force is considered. Note that for the
velocity dependent friction model the overall COF is different from the local COF at a
node. For Coulomb’s law with a constant COF, in contrast, fr equals the local COF.
COF is not an intrinsic property of materials (Ludema, 2001). Besides the
sliding speed, COF also varies with a great variety of factors like contact pressure,
surface lubrication or contamination, roughness, temperature, humidity, wear, etc.
(Ludema, 2001; Bucher et al., 2006; Blau, 2009; Arias-Cuevas et al., 2010a, b). In a twin-
disc rig, the traction coefficient corresponding to a constant creepage of 1.5 percent was
found to decrease from 0.436 to 0.329 under dry-clean conditions as the circumferential
speed increased from 0.5 to 2 m/s (Deters and Proksch, 2005). Because the friction
saturated at a creepage of about 1.0 percent for the tested cases, the above-mentioned
traction coefficient decrease, to a large extent, revealed the dependence of COF on
velocity. Such a result is quite close to the velocity dependence of the COF for a steel-on-
steel sliding contact measured in Sampson et al. (1943), being from 0.44 to 0.24 with the
increase of sliding velocity. Further considering that the wheel-rail COF under the dry-
clean condition, treated in this paper, is reported to be typically 0.4-0.65 (Cann, 2006),
and the measurement of COF is out of the scope of this work, the authors assume the
static and kinematic COFs as 0.5 and 0.32, respectively, with a certain randomness, as
listed in Table I. The exponential decay coefficient k of the COF model, which defines
the velocity dependence of COF plotted in Figure 2, is determined by fitting the COF
results measured in Sampson et al. (1943). Note that the friction model shown in
Figure 2 does not refer to a specific test result, but represents the typical values under
the dry-clean condition. This is a compromise of the following fact: for the wheel-rail
contact, as an open system, the COF in between can vary significantly when different
lubricant is present (Cann, 2006). Recently, on the basis of field tests Allotta et al. (2014)
and Meli and Ridolfi (2015) developed an efficient model to introduce the dependence of
the contaminated COF on the energy dissipation into the multi-body simulation.
Material damping, which dissipates the energy stored in continuum vibrations, is
considered by the Rayleigh damping (Cm) that is generally expressed as:
C m ¼ aM þ bK m (3)
where α and β are the mass (M) and stiffness (Km) proportional damping
constants, respectively. In this work, mass proportional damping is ignored
EC
0.50
33,4

0.45

Local COF
1038 0.40

0.35

Figure 2.
Simulated
dependence of local 0.30
0.0 0.2 0.4 0.6 0.8 1.0
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

COF on micro-slip
Micro-slip, sn (m/s)

because it is effective for low-frequency vibrations (Kazymyrovych et al., 2010). Value


of β is taken based on estimate made by inverse modeling (Kazymyrovych
et al., 2010).

3. Results
3.1 With traction coefficient of 0.3
Two different friction models, one is Coulomb’s law of friction with a constant
COF (f ¼ fS) and the other defined by Equation (1), are employed to simulate the
same case with the traction coefficient being 0.3. It is first found that no considerable
change is made to the normal contact force when the velocity dependence of COF is
included, and the resulted variations of the longitudinal contact force are negligible
(less than 0.5 percent). Note that in the simulations, a constant driving torque is
applied to the wheel axis (valid for all simulations of this work), while the transmitted
longitudinal force may vary because of vibrations of the system excited by the
moving wheel.
Contact patch and contact stresses. Figure 3 shows the nodes in contact when different
friction models are employed. A node is considered to be in contact if constraint (4) is met:

A node is in contact if : F n_N 40 (4)

where Fn_N is the nodal force in the direction normal to the local surface. The black blocks
in Figure 3 indicate nodes where the solutions of the two friction models differ. It is seen
that the difference occurs only at the border of the contact patch where the contact pressure
is negligibly small. The corresponding distributions of pressure ( p) and surface shear stress
(τ) along the longitudinal axis of the contact patch (y ¼ 0, it is referred to as the longitudinal
axis hereinafter) are given in Figure 4.
It is observed from Figures 3 and 4 that the normal contact solution, namely, the shape
and size of the contact patch and pressure distribution, is not affected by the friction
models. In contrast, the surface shear stress distribution significantly changes with the
friction model. In the trailing part of the contact patch where the slip area is located at, the
surface shear stress reduces as the velocity dependence of COF is included. This is because
the local COF becomes smaller than fS when micro-slip occurs. To achieve the prescribed
Common nodes for both friction models
Discrepant nodes
Explicit finite
6.0 element
Lateral, y (mm) method
3.0

0.0 1039

–3.0
Figure 3.
Nodes in contact
–6.0 Rolling direction when different
friction models
–8.0 –4.0 0.0 4.0 8.0
are applied
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

Longitudinal, x (mm)

1,000
Rolling direction f varies with v
p × fS
800 τ
Constant fS
p × fS
Stress (MPa)

600 τ

Figure 4.
400 Pressure ( p) and
surface shear stress
(τ) distributions
200
along the
longitudinal axis
0 (y ¼ 0.0) when
different friction
–8.0 –4.0 0.0 4.0 8.0
models are applied
Longitudinal, x (mm)

traction coefficient the surface shear stress is inevitably increased in the leading part of the
contact patch. From now on, analyses only focus on the tangential contact solution.
Areas of adhesion and slip. Distinction between areas of adhesion and slip was made
based on nodal forces by inequality (5) in Zhao and Li (2011):
A node is in adhesion if : f jF nN jjF nT j 4eT (5)
where f is a constant COF (no velocity dependence was considered), Fn_T is the nodal
force in the tangential plane and εT is a tolerance. εT was taken as 0.3 percent of the
maximal tangential nodal force in the contact patch (Zhao and Li, 2011).
In this work, the criterion of inequality (5) becomes not applicable because local COF
changes from node to node due to its dependence on micro-slip. Instead, another
criterion, as given by inequality (6), is employed to distinguish the slip area from the
adhesion one:
A node is in slip if : sn 4 es (6)
EC where sn is the micro-slip at a node, and εs is a tolerance. To determine an appropriate
33,4 value of εs, the distinction between adhesion and slip areas is judged by both
inequalities (5) and (6) for the case with constant friction coefficient. It is found that
εs ¼ 0 is applicable for the discretized system and a value of 0.05 m/s is suitable for εs,
as shown in Figure 5. Such a value is 12.5 percent of the maximum micro-slip in the
current case (the maximum is about 0.4 m/s), being acceptable for engineering
1040 purposes. The essence of εs may be errors related to the transient FE modeling, which
needs further investigation to confirm. The un-sharp distinction between slip and
adhesion areas in Figure 5(a) and the odd node in slip in the middle of the adhesion area
in Figure 5(b) should be owing to the numerical errors and the criterion of inequality (6).
A better criterion than inequality (6) is needed if more accurate distinction is required.
With the new criterion, the adhesion-slip distinction is obtained from micro-slip
results for the case with velocity dependent COF, as shown in Figure 6. It is very
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

different from that with constant COF in Figure 5. Furthermore, the distinction seems
not in agreement with the surface shear stress distribution given in Figure 4.
The essence of these discrepancies is that steady state rolling contact is not
achieved in the case with the velocity dependent COF, unlike with constant COF.

(a) (b)
Node in slip Node in slip
Rolling direction Node in adhesion Rolling direction Node in adhesion
6.0 6.0
Lateral, y (mm)

Lateral, y (mm)

3.0 3.0

0.0 0.0

–3.0 –3.0

Figure 5.
–6.0 –6.0
Distinction of slip
and adhesion areas –8.0 –4.0 0.0 4.0 8.0 –8.0 –4.0 0.0 4.0 8.0
in the case with Longitudinal, x (mm) Longitudinal, x (mm)
constant COF
Notes: (a) By inequality (5) as in Zhao and Li (2011); (b) by inequality (6) with s = 0.05 m/s

Node in slip
Rolling direction
Node in adhesion
6.0
Lateral, y (mm)

3.0

0.0
Figure 6.
Distinction of slip –3.0
and adhesion
areas in the case
with velocity –6.0
dependent friction
–8.0 –4.0 0.0 4.0 8.0
by inequality (6)
Longitudinal, x (mm)
This is why a case with constant COF is employed above to determine the tolerance Explicit finite
εs. More detailed explanations of the un-steady state rolling contact are given in element
the next section by examining the oscillations of tangential contact related to the
velocity dependence.
method
Oscillations of tangential contact. Besides the distinction between adhesion and slip
areas at instant t (Figure 6), results at another two instants are also obtained, namely, at
t − 0.0057 and t + 0.0057 ms, as shown in Figure 7. It is seen that the areas of slip and 1041
adhesion change greatly with time, especially in the central zone of the contact patch.
Note that in post procession, limited by computer capacity, the sample frequency is taken
as 175 kHz (i.e. 1/0.0057 ms), which is why the results at t − 0.0057 and t + 0.0057 ms
are presented in Figure 7.
To show the reason behind such a phenomenon, velocities of two nodes are plotted
vs the longitudinal position of the wheel center (hereinafter, it is referred to as wheel
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

position) in Figure 8. The two nodes are located on the longitudinal axis during contact,
one of which is in the rail surface and the other in the wheel surface. They enter and
leave the contact patch when the wheel is located at about x ¼ −8.0 and 8.0 mm,
respectively. Figure 9 shows the corresponding results with constant COF. Comparing
Figures 8 and 9, it is found that oscillations of very high frequencies are caused by the

(a) (b)
Node in slip Node in slip
Rolling direction Rolling direction Node in adhesion
Node in adhesion
6.0 6.0
Lateral, y (mm)
Lateral, y (mm)

3.0 3.0

0.0 0.0
Figure 7.
–3.0 –3.0 Distinction of slip
and adhesion areas
–6.0 –6.0 at another two
–8.0 –4.0 0.0 4.0 8.0 –8.0 –4.0 0.0 4.0 8.0 instants in the case
Longitudinal, x (mm) Longitudinal, x (mm) with velocity
dependent friction
Notes: (a) t−Δt; (b) t+Δt. Δt = 0.0057 ms

(a) (b)
0.8 0.8
A node in rail surface A node in wheel surface
0.6 x 0.6 x
y y
0.4 z 0.4 z
Velocity (m/s)
Velocity (m/s)

0.2 0.2

0.0 0.0

–0.2 –0.2

–0.4 –0.4
Figure 8.
–0.6 –0.6 Velocities of nodes in
–8.0 –4.0 0.0 4.0 8.0 12.0 –8.0 –4.0 0.0 4.0 8.0 12.0 the case with
Wheel position, x (mm) Wheel position, x (mm) velocity dependent
friction
Notes: (a) At rail top; (b) in wheel contact surface
EC (a) (b)
0.8 0.8
33,4 0.6
The node in rail surface
x 0.6
The node in wheel surface
x
y y
0.4 z 0.4 z
Velocity (m/s)

Velocity (m/s)
0.2 0.2

0.0 0.0
1042 –0.2 –0.2

–0.4 –0.4

Figure 9. –0.6 –0.6


Velocities of nodes –8.0 –4.0 0.0 4.0 8.0 12.0 –8.0 –4.0 0.0 4.0 8.0 12.0
in the case with Wheel position, x (mm) Wheel position, x (mm)
constant COF
Notes: (a) At rail top; (b) in wheel contact surface
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

velocity dependence of COF. Applying a FFT, it is found that the very high-frequency
oscillations correspond to the vibration components in the frequency range between 40
and 50 kHz. This explains the great changes of the stick-slip distinction with time
shown in Figures 6 and 7. To interpret the results in Figures 6 and 7, it should also be
born in mind that the oscillations are very local phenomena since the COF varies from
node to node depending on the micro-slip, and the oscillations of different nodes may
affect each other significantly because in the period of 0.0057 ms the generated waves
only propagates 33.7 mm (the sonic speed is 5920 m/s in steel).
Note that the two nodes presented in Figure 8, when they are in the contact patch,
are not always in contact with each other due to the existence of micro-slip. Hence, the
velocities of the two nodes do not synchronize in Figure 8. It should be noted that it is
not yet confirmed if the stick-slip phenomenon related to the velocity dependent COF is
realistic. It could contain both a realistic part which may be caused by high-frequency
continuum vibrations of the wheel and rail in the contact excited by the moving
wheel and the velocity dependency of COF, and a pseudo part which is owing to
numerical errors.
It is further observed from Figure 8 that the magnitudes of the variations of nodal
velocities are relatively small at the beginning, i.e. when the nodes are in the leading
part of the contact patch. Gradually, when they approach and enter the trailing part
where a slip area is expected for a constant COF, larger oscillations occur. Such a
phenomenon is explained by the fact that a relatively stable state, i.e. definitely in
adhesion, exists in the leading part, as shown in Figures 6 and 8. In other words,
oscillations caused by the velocity dependent COF mainly occur in the trailing part
where micro-slip exists. Velocity variations in the leading part should be due to
transmission of the excited oscillations.
Figure 10 shows distributions of micro-slip (magnitude) along the longitudinal axis
at different instants when velocity dependence of COF is considered. Results of the
same instants as in Figures 6 and 7 are presented. It is seen that the micro-slip
varies greatly, even though the time difference is so tiny between the three instants that
the contact patches almost coincide in space. This is in line with the nodal velocities
given in Figure 8. Note that a result of constant COF is also plotted in Figure 10
for comparison.
Unlike micro-slip, the surface shear stress distribution is smooth and does not
change considerably at the selected instants. This is in agreement with the negligible
1.0 Explicit finite
Rolling direction Instant t -Δt
Instant t
element
Micro-slip, sn (m/s)
0.8 Instant t +Δt method
Constant f
0.6

1043
0.4

0.2
Figure 10.
Micro-slip
0.0 distributions along
the longitudinal axis
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

–8.0 –4.0 0.0 4.0 8.0


at different instants
Longitudinal, x (mm)

influence of the velocity dependence of COF on the resultant friction force. Such a
phenomenon may be explained by the fact that the stress is obtained from strain
calculated from nodal displacement of the elements. Because the displacement is
calculated by time integration of the nodal velocity, the integration filters out the high-
frequency variation in the velocity, leading to the phenomenon that the stress
distributions are smooth and almost constant.
Distributions of pressure and surface shear stress are shown in Figure 11, from
which it can be seen that the local COF approaches the kinematic one in the trailing part
where a slip area is expected for a constant COF. To explain this phenomenon, the
average micro-slip in the trailing part of the contact patch is plotted vs wheel position
in Figure 12. It is observed that at instant t (i.e. at wheel position x ¼ 0) the average
micro-slip is 0.29 m/s, which in Figure 2 corresponds to a COF of 0.35 being close to the
kinematic COF (fD ¼ 0.32). Note that the average micro-slip shown in Figure 12 is
calculated among all the nodes in the trailing part of the contact patch which is defined

800
Rolling direction At instant t
P×fD
τ
600 P×fS
Stress (MPa)

400

Figure 11.
Comparison between
200 pressure and surface
shear stress along
the longitudinal axis
0 in the case with
velocity dependent
–8.0 –4.0 0.0 4.0 8.0
friction
Longitudinal, x (mm)
EC 0.8
Original results
33,4 Rolling direction
Smoothed results

Average micro-slip, sn (m/s)


0.6

1044 0.4

0.2

0.0
Figure 12.
–6.0 –3.0 0.0 3.0 6.0
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

Variation of the
average micro-slip in Wheel position, x (mm)
the trailing part of
the contact patch Note: Defined by x < 4.0 mm for instant t and changes
correspondingly with the movement of the contact patch

by x o 4.0 mm for instant t (see Figure 6) and changes correspondingly with the
movement of the contact patch.
It has to be clarified that the local COF is determined by the micro-slip at a node, not
by the average level given in Figure 12. Therefore, micro-slip variations with nodes at
an instant (see Figure 10) should also be considered to understand Figure 11. Moreover,
Figure 12 also shows that the average micro-slip significantly varies with time, being in
line with results shown above.
Because of its influence on the surface shear stress, the velocity dependence of COF
greatly modifies the Von Mises (V-M) stress distribution. Due to the space limit, results
of V-M stresses are not presented here.

3.2 Different traction coefficients


By specifying the corresponding driving torque applied to the wheel axis, traction
coefficient is varied to investigate the influence of velocity dependent COF under
different friction exploitation levels. In later statements, cases with different driving
torques are distinguished from each other by their nominal traction coefficients
(referred to as μn). A nominal traction coefficient is approximately proportional to the
driving torque, which is valid if an infinitely large COF exists. For the case with a finite
COF, when the friction is unsaturated (i.e. μn o fD), the nominal traction coefficient is
the same as the actually achieved traction coefficient (referred to as μa), e.g. in the case
discussed in last section. Once saturation occurs, however, further increase of driving
torque (or the nominal traction coefficient) will not increase the actual traction
coefficient (limited by the overall COF fr).
When traction coefficient is 0.5, changes of the longitudinal contact force caused by
the velocity dependence of COF are plotted in Figure 13, which are obtained by
subtracting the longitudinal force of constant COF from that of velocity dependent
COF. It is seen that, unlike the case discussed in last section, influence of velocity
dependent COF becomes significant as the friction saturates. Such a phenomenon is as
expected because the actually achieved traction coefficient is the same as the nominal
one in the case with constant COF, but not in the case with velocity dependent COF
0 Explicit finite

Changes of longitudinal contact force (kN)


element
method
–10

–20 1045

–30 13 mm Figure 13.


Changes of the
longitudinal contact
force caused by the
–40 velocity dependence
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

–10.0 –5.0 0.0 5.0 10.0


of COF when μ ¼ 0.5
Wheel position, x (mm)

(the overall COF approaches the kinematic COF). The variation with a wavelength of
13 mm (ΔFL1), as indicated in Figure 13, is due to a variation of the normal contact force
(ΔP) related to an eigen mode of the vertical wheel-rail interaction:

DF L1 ¼ DP ma mn (7)

Note that the variation of ΔFL1 vanishes if friction is unsaturated, i.e. μa ¼ μn.
Figure 14 shows variations of the longitudinal velocity of the chosen rail surface
node under different traction coefficients. For easy comparison, different vertical shifts
are applied to the cases in the figure, i.e. the lines of zero velocity in the figure do not
coincide with each other for different cases. It is seen that with the increase of traction
coefficient the velocity variation starts earlier due to the shrinkage of adhesion area.
Meanwhile, level of the variation (or of the vibration in behind) increases with the
traction coefficient, especially after friction saturation.

μ = 0.1
Longitudinal velocity

μ = 0.2

μ = 0.3

1 m/s
μ = 0.4

Figure 14.
μ = 0.5 Longitudinal velocity
variations of the
chosen rail surface
node under different
–10.0 –5.0 0.0 5.0 10.0
traction coefficients
Wheel position, x (mm)
EC With the increase of traction coefficient, the surface shear stress distribution along the
33,4 longitudinal axis changes as shown in Figure 15. The results are in agreement with
the classic solutions of rolling contact. Furthermore, it is seen that as traction coefficient
changes from 0.1 to 0.3 the surface shear stress distribution is almost unchanged in the
trailing part of the contact patch. This is because among these cases the local COF has
almost the same value in the trailing part, determined by their similar micro-slip levels
1046 (see the similar vibration levels in Figure 14). In contrast, the local COF in the trailing
part of the contact patch reduces considerably when friction saturates, i.e. the traction
coefficient increases from 0.3 to 0.4. Note that the normal contact solution is the
same among these cases because it is not affected by the friction models and
the traction coefficient. The result of μ ¼ 0.5 is not given in Figure 15 because it is close
to that of μ ¼ 0.4.
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

4. Discussions
4.1 About contact solutions
In studies of sliding contact, it has been found that the variation of COF in static and
kinematic states and the velocity dependence of kinematic COF can lead to the so called
non-uniform sliding or tangential oscillations (Adams, 1998; Ibrahim, 1994). Such
instability may lead to squealing noise in practice (Nouby et al., 2011). Results
presented above, e.g. Figures 6 and 7, show that such a friction-induced stick-slip
phenomenon also exists in the trailing part of rolling contact patch.
It has been found that, for a plane contact patch existing between bodies of the same
linear-elastic material, the friction-induced oscillations have no considerable influence
on the normal contact solution. In other words, for the studied cases the normal
contact problem can still be decoupled from the tangential one when tangential
contact oscillates.
The very high-frequency oscillations caused by the velocity dependent COF
are self-excited in nature, so that they do not die away with time and lead to
continuous oscillation of the tangential contact. It is found in this work that the micro-
slip greatly varies with time as the dependence is considered, while the tangential
contact stress is quite stable. Therefore, friction models with constant kinematic
COFs, if defined with an appropriate value, may be sufficient for stress estimate.

600
μ = 0.1
Rolling direction μ = 0.2
Surface shear stress (MPa)

μ = 0.3
μ = 0.4

400

200
Figure 15.
Surface shear stress
distributions along
the longitudinal axis
under different 0
–8.0 –4.0 0.0 4.0 8.0
traction coefficients
Longitudinal, x (mm)
However, for studies into damages such as wear, which are related to the Explicit finite
tangential stress and the micro-slip, more sophisticated friction models may have to element
be employed. To capture the micro-slip variations caused by the velocity dependent
COF, very fine time steps are required due to the high-frequency characteristic.
method
For example, the time step is 2.6 × 10−8 s for the model developed in this work as
mentioned above.
1047
4.2 About modeling
In the velocity dependent friction model, the static and the kinematic COFs are
assumed constant and the COF monotonically changes in a low-velocity range, i.e.
mainly less than about 0.5 m/s. This is taken based on the test results reported in the
literature as mentioned above. In field, all these parameters may change significantly
under different conditions, causing considerable influence to the instability problems
such as squealing noise. Thus, more complicated or different velocity dependence may
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

need to be considered for various engineering problems.


Different velocity dependence of COF has also been considered by changing the
value of the exponential decay coefficient k to examine its influence. It is confirmed
that phenomena explained above exist for all cases, but with some magnitude
difference. It should be noted that errors of the FE method may also in the simulations
contribute to the oscillations excited by the velocity dependence. This is important
because the oscillations are very local phenomena, occurring in a territory of a few
elements as shown in Figure 10. In other words, it is still unknown to which extent the
simulated oscillations are physical. Hence, further investigations are needed to better
understand the phenomenon, including the influence of different friction dependence.
To such an end, the authors propose to first use more powerful computers to take into
account all results in post procession. Furthermore, by varying the material and
friction parameters, the relative magnitude of the physical behavior with respect to
the numerical error can be changed, through which the physics and numerical error
can be distinguished to a certain extent. In addition, the interpolation method
employed to calculate the micro-slip should also be optimized from the aspect of
numerical error. It might be worthwhile to mention that an experimental validation of
the proposed approach remains a great challenge since there are no effective
experimental techniques today for the determination of contact stress and
strain states, especially under dynamic conditions. In the near future, however, the
resultant contact forces obtained from the approach can be validated against other
strategies (Polach, 2005; Giménez et al., 2005; Xie et al., 2006; Zhang et al., 2002;
Arias-Cuevas et al., 2010a, b).
With the presented approach, the traction behavior of a wheel is directly related to
the contact stresses due to detailed calculations of contact with a sophisticated
friction model. Such an approach may further be employed for detailed investigations
into the traction behavior of vehicles under various contact surface conditions
(e.g. dry-clean, contaminated, etc.) and the related damages. Effects of friction
modifiers on rail damages may also be studied with it for different rolling speeds.
At last, due to its high computation costs, such an approach is hard to be introduced
into the multi-body models of railway vehicles (e.g. Shabana et al., 2004; Pombo
and Ambrosio, 2008), whereas it may be employed as a potential calibration and
validation tool for the low-computation cost approaches like the one developed by
Allotta et al. (2014), or to make tables or curve-fit empirical formula that can be
embedded in multi-body models.
EC 5. Conclusions
33,4 A velocity dependent friction model is introduced into the explicit FE model of wheel-
rail rolling contact to examine its influence on detailed contact solutions. By comparing
to the results of a constant COF, the following conclusions can be drawn from elastic
simulations:
(1) Very high-frequency oscillations are caused in the tangential plane by the
1048 dependence. They occur in the trailing part of the contact patch where micro-
slip exists, and can be transmitted to the leading part. As a result, the micro-slip
distribution greatly varies with time. However, the surface shear stress
distribution is still stable.
(2) The dependence has no influence on the normal contact force and the pressure.
The normal contact can still be decoupled from the tangential one when
tangential contact oscillates, if the rolling elements are made of the same linear-
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

elastic material.
(3) When friction is unsaturated, negligible influence is caused to the tangential
contact force by the dependence. In contrast, the surface shear stress
distribution significantly changes with the employed friction model, because the
local COF changes correspondingly, especially in the trailing part of the contact
patch where slip area is located at.
(4) Once friction saturates, the tangential contact force can be greatly influenced by
the dependence because the overall COF approaches static COF fs in the case
with constant friction model and equates kinematic COF fD in the case including
the velocity dependence.
(5) With the velocity dependent friction, magnitudes of the excited vibrations
slightly increase with the traction coefficient before friction saturation. After
saturation, the vibrations become obviously fiercer.
(6) The maximum surface shear stress increases with traction coefficient before
friction saturation, and slightly decreases after that.
(7) The criterion of inequality (6) is acceptable for engineering purposes to
distinguish the adhesion and slip areas from each other. The essence of εs needs
to be further investigated and a better criterion is preferred if more accurate
distinction is required.
(8) The non-uniform sliding or tangential oscillations observed in sliding contact
also exist in the slip area of rolling contact.

References
Adams, G.G. (1998), “Steady sliding of two elastic half-spaces with friction reduction due to
interface stick-slip”, Journal of Applied Mechanics: ASME, Vol. 65 No. 2, pp. 470-475.
Allotta, B., Meli, E., Ridolfi, A. and Rindi, A. (2014), “Development of an innovative wheel-rail
contact model for the analysis of degraded adhesion in railway systems”, Tribology
International, Vol. 69, January, pp. 128-140.
Arias-Cuevas, O., Li, Z., Lewis, R. and Gallardo-Hernandez, E.A. (2010a), “Rolling-sliding
laboratory tests of fiction modifiers in dry and wet wheel-rail contacts”, Wear, Vol. 268
Nos 3-4, pp. 543-551.
Arias-Cuevas, O., Li, Z., Popovici, R.I. and Schipper, D.J. (2010b), “Simulation of curving Explicit finite
behaviour under high traction in lubricated wheel-rail contacts”, Vehicle System Dynamics,
Vol. 48 No. S1, pp. 299-316.
element
method
Benson, D.J. and Hallquist, J.O. (1990), “A single surface contact algorithm for the post-buckling
analysis of shell structures”, Computer Methods in Applied Mechanics and Engineering,
Vol. 78 No. 2, pp. 141-163.
Blau, P.J. (2009), “Embedding wear models into friction models”, Tribology Letters, Vol. 34 No. 1, 1049
pp. 75-79.
Bucher, F., Dmitriev, A.I., Ertz, M., Knothe, K., Popov, V.L., Psakhie, S.G. and Shilko, E.V. (2006),
“Multiscale simulation of dry friction in wheel/rail contact”, Wear, Vol. 261 Nos 7-8,
pp. 874-884.
Cann, P.M. (2006), “‘Leaves on the line’ problem – a study of leaf residue film formation and
lubricity under laboratory test conditions”, Tribology Letters, Vol. 24 No. 2, pp. 151-158.
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

Courant, R., Friedrichs, K. and Lewy, H. (1928), “On the partial difference equations of
mathematical physics”, Mathematische Annalen, Vol. 100, pp. 32-74.
Deters, L. and Proksch, M. (2005), “Friction and wear testing of rail and wheel material”,
Wear, Vol. 258 Nos 7-8, pp. 981-991.
Dollevoet, R., Li, Z. and Arias-Cuevas, O. (2010), “A method for the prediction of head checking
initiation location and orientation under operational loading conditions”, Journal of Rail
and Rapid Transit, Vol. 224 No. 5, pp. 369-374.
Giménez, J.G., Alonso, A. and Gómez, E. (2005), “Introduction of a friction coefficient dependent
on the slip in the fastsim algorithm”, Vehicle System Dynamics, Vol. 43 No. 4, pp. 233-244.
Goryacheva, I.G., Rajeev, P.T. and Farris, T.N. (2001), “Wear in partial slip contact”, Journal of
Tribology, Vol. 123 No. 4, pp. 848-856.
Ibrahim, R.A. (1994), “Friction-induced vibration, chatter, squeal, and chaos: part ii – dynamics
and modeling”, ASME Applied Mechanics Reviews, Vol. 47 No. 7, pp. 227-253.
Kalker, J.J. (1990), Three-Dimensional Elastic Bodies in Rolling Contact, Kluwer Academic
Publishers, Dordrecht.
Kazymyrovych, V., Bergström, J. and Thuvander, F. (2010), “Local stresses and material damping
in very high cycle fatigue”, International Journal of Fatigue, Vol. 32 No. 10, pp. 1669-1674.
Lance, B.J. and Sadeghi, F. (1993), “The normal approach and stick-slip phenomena at the
interface of two rough bodies”, Journal of Tribology, Vol. 115 No. 3, pp. 445-452.
Li, Z. and Kalker, J.J. (1998), “Simulation of severe wheel-rail wear”, Proceedings of the Sixth
International Conference on Computer Aided Design, Manufacture and Operation in the
Railway and Other Mass Transit Systems, Lisbon, September, pp. 393-402.
Li, Z., Zhao, X., Esveld, C., Dollevoet, R. and Molodova, M. (2008), “An investigation into the
causes of squats: correlation analysis and numerical modeling”, Wear, Vol. 265 Nos 9-10,
pp. 1349-1355.
Ludema, K.C. (2001), “Friction”, in Bhushan, B. (Ed.), Modern Tribology Handbook, Chapter 5,
CRC Press LLC, New York, NY.
Lugt, P.M. (2009), “A review on grease lubrication in rolling bearings”, Tribology Transactions,
Vol. 52 No. 4, pp. 470-480.
Meli, E. and Ridolfi, A. (2015), “An innovative wheel–rail contact model for railway vehicles under
degraded adhesion conditions”, Multibody System Dynamics, Vol. 33 No. 3, pp. 285-313.
Nouby, M., Abdo, J., Mathivanan, D. and Srinivasan, K. (2011), “Evaluation of disc brake
materials for squeal reduction”, Tribology Transactions, Vol. 54 No. 4, pp. 644-656.
EC Polach, O. (2005), “Creep forces in simulations of traction vehicles running on adhesion limit”,
Wear, Vol. 258 Nos 7-8, pp. 992-1000.
33,4
Pombo, J. and Ambrosio, J. (2008), “Application of a wheel-rail contact model to railway dynamics
in small radius curved tracks”, Multibody System Dynamics, Vol. 19 No. 1, pp. 91-114.
Popescu, G., Morales-Espejel, G.E., Wemekamp, B. and Gabelli, A. (2006), “An engineering model
for three-dimensional elastic-plastic rolling contact analyses”, Tribology Transactions,
1050 Vol. 49 No. 3, pp. 387-399.
Sadeghi, F., Jalalahmadi, B., Trevor, S.S., Raje, N. and Arakere, N.K. (2009), “A review of rolling
contact fatigue”, Journal of Tribology, Vol. 131 No. 4, pp. 1-15.
Sampson, J.B., Morgan, F., Reed, D.W. and Muskat, M. (1943), “Studies in lubrication: xii. friction
behavior during the slip portion of the stick slip process”, Journal of Applied Physics,
Vol. 14 No. 12, pp. 689-700.
Shabana, A., Zaazaa, K.E., Escalona, J.L. and Sanyc, J.R. (2004), “Development of elastic
force model for wheel/rail contact problems”, Journal of Sound and Vibration, Vol. 269
Downloaded by West Virginia University At 01:30 07 June 2016 (PT)

Nos 1-2, pp. 295-325.


Vollebregt, E.A.H. and Schuttelaars, H.M. (2012), “Quasi-static analysis of two-dimensional
rolling contact with slip-velocity dependent friction”, Journal of Sound and Vibration,
Vol. 331 No. 9, pp. 2141-2155.
Warhadpande, A., Sadeghi, F., Evans, R.D. and Kotzalas, M.N. (2012), “Influence of plasticity-
induced residual stresses on rolling contact fatigue”, Tribology Transaction, Vol. 55 No. 4,
pp. 422-437.
Xie, G., Allen, P.D., Iwnicki, S.D., Alonso, A., Thompson, D.J., Jones, C.J.C. and Huang, Z.Y. (2006),
“Introduction of falling friction coefficients into curving calculations for studying curve
squeal noise”, Vehicle System Dynamics, Vol. 44, Supplement, pp. 261-271.
Xu, B. and Jiang, Y. (2002), “Elastic-plastic finite element analysis of partial slip of rolling
contact”, Journal of Tribology, Vol. 124 No. 1, pp. 20-26.
Zhang, J., Sun, S. and Jin, X. (2009), “numerical simulation of two-point contact between wheel and
rail”, Acta Mechanica Solida Sinica, Vol. 22 No. 4, pp. 352-359.
Zhang, W., Chen, J., Wu, X. and Jin, X. (2002), “Wheel/rail adhesion and analysis by using full
scale roller rig”, Wear, Vol. 253 Nos 1-2, pp. 82-88.
Zhao, X. and Li, Z. (2011), “The solution of frictional wheel-rail rolling contact with a 3-d transient
finite element model: validation and error analysis”, Wear, Vol. 271 Nos 1-2, pp. 444-452.
Zhao, X. and Li, Z. (2015), “A 3-D finite element solution of frictional wheel–rail rolling contact in
elasto-plasticity”, Journal of Engineering Tribology, Vol. 229 No. 1, pp. 86-100.

Corresponding author
Xin Zhao can be contacted at: xinzhao@home.swjtu.edu.cn

For instructions on how to order reprints of this article, please visit our website:
www.emeraldgrouppublishing.com/licensing/reprints.htm
Or contact us for further details: permissions@emeraldinsight.com

You might also like