12
ORIGINAL PAPER Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian/Generalized Newtonian Lubricant W. Habchi D. Eyheramendy S. Bair P. Vergne G. Morales-Espejel Received: 4 December 2007 / Accepted: 25 February 2008 / Published online: 8 March 2008 Ó Springer Science+Business Media, LLC 2008 Abstract The classical ElastoHydroDynamic (EHD) theory assumes a Newtonian lubricant and an isothermal operating regime. In reality, lubricating oils do not behave as perfect Newtonian fluids. Moreover, in most operating conditions of an engineering system, especially at high speeds, thermal effects are important and temperature can no longer be considered as constant throughout the system. This is one reason why there has always been a gap between numerical results and experimental data. This paper aims to show that this gap can be reduced by taking into consideration the heat generation that takes place in the contact and using appropriate rheological models. For this, a unique thermal ElastoHydrodynamic lubrication model is developed for both Newtonian and non-Newto- nian lubricants. Pressure, film thickness and traction results are then compared to their equivalent isothermal results and experimental data. The agreement between thermal calculations and experiments reveals the necessity of con- sidering thermal effects in EHD models. Keywords EHL with non-Newtonian lubricants Thermal effects in EHL Nomenclature q Lubricant’s density g Generalized Newtonian viscosity l Newtonian viscosity H Film thickness p Pressure p h Hertzian contact pressure a Hertzian contact radius x,y,z Space coordinates U m Mean entrainment velocity U Elastic displacement vector (U = {u, v, w}) L Load u i Surface velocity of body i u f , v f Fluid flow velocity components in the x- and y-directions, respectively h 0 Film thickness equation constant R Ball’s radius r Stress tensor e Strain tensor C Compliance matrix T Temperature T 0 Ambient temperature T R Reference temperature c i Heat capacity of body i q i Density of body i k i Thermal conductivity of body i E i Young’s modulus of body i t i Poisson’s ratio of body i x in Inlet abscissa of the contact p + Positive part of the pressure distribution l 0 Lubricant’s zero pressure Newtonian viscosity W. Habchi (&) D. Eyheramendy P. Vergne LaMCoS, INSA-Lyon, CNRS UMR5259, Lyon 69621, France e-mail: [email protected] D. Eyheramendy Ecole Centrale Marseille, Laboratoire de Me ´canique et d’Acoustique, 13451 Marseille Cedex 20, France S. Bair G.W. Woodruff School of Mechanical Engineering, Centre for High-Pressure Rheology, Georgia Institute of Technology, Atlanta, GA 30332-0405, USA G. Morales-Espejel SKF Engineering and Research Center, Nieuwegein, The Netherlands 123 Tribol Lett (2008) 30:41–52 DOI 10.1007/s11249-008-9310-9

Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

  • Upload
    vishara

  • View
    107

  • Download
    3

Embed Size (px)

Citation preview

Page 1: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

ORIGINAL PAPER

Thermal Elastohydrodynamic Lubrication of Point ContactsUsing a Newtonian/Generalized Newtonian Lubricant

W. Habchi Æ D. Eyheramendy Æ S. Bair ÆP. Vergne Æ G. Morales-Espejel

Received: 4 December 2007 / Accepted: 25 February 2008 / Published online: 8 March 2008

Springer Science+Business Media, LLC 2008

Abstract The classical ElastoHydroDynamic (EHD)

theory assumes a Newtonian lubricant and an isothermal

operating regime. In reality, lubricating oils do not behave

as perfect Newtonian fluids. Moreover, in most operating

conditions of an engineering system, especially at high

speeds, thermal effects are important and temperature can

no longer be considered as constant throughout the system.

This is one reason why there has always been a gap

between numerical results and experimental data. This

paper aims to show that this gap can be reduced by taking

into consideration the heat generation that takes place in

the contact and using appropriate rheological models. For

this, a unique thermal ElastoHydrodynamic lubrication

model is developed for both Newtonian and non-Newto-

nian lubricants. Pressure, film thickness and traction results

are then compared to their equivalent isothermal results

and experimental data. The agreement between thermal

calculations and experiments reveals the necessity of con-

sidering thermal effects in EHD models.

Keywords EHL with non-Newtonian lubricants Thermal effects in EHL

Nomenclature

q Lubricant’s density

g Generalized Newtonian viscosity

l Newtonian viscosity

H Film thickness

p Pressure

ph Hertzian contact pressure

a Hertzian contact radius

x,y,z Space coordinates

Um Mean entrainment velocity

U Elastic displacement vector (U = u, v, w)

L Load

ui Surface velocity of body i

uf, vf Fluid flow velocity components in the x- and

y-directions, respectively

h0 Film thickness equation constant

R Ball’s radius

r Stress tensor

e Strain tensor

C Compliance matrix

T Temperature

T0 Ambient temperature

TR Reference temperature

ci Heat capacity of body i

qi Density of body i

ki Thermal conductivity of body i

Ei Young’s modulus of body i

ti Poisson’s ratio of body i

xin Inlet abscissa of the contact

p+ Positive part of the pressure distribution

l0 Lubricant’s zero pressure Newtonian

viscosity

W. Habchi (&) D. Eyheramendy P. Vergne

LaMCoS, INSA-Lyon, CNRS UMR5259, Lyon 69621, France

e-mail: [email protected]

D. Eyheramendy

Ecole Centrale Marseille, Laboratoire de Mecanique

et d’Acoustique, 13451 Marseille Cedex 20, France

S. Bair

G.W. Woodruff School of Mechanical Engineering,

Centre for High-Pressure Rheology, Georgia Institute

of Technology, Atlanta, GA 30332-0405, USA

G. Morales-Espejel

SKF Engineering and Research Center, Nieuwegein,

The Netherlands

123

Tribol Lett (2008) 30:41–52

DOI 10.1007/s11249-008-9310-9

Page 2: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

q0 Lubricant’s zero pressure density

SRR Slide-to-roll ratio = 2(us - up)/(us + up)

Hc Central film thickness

Hmin Minimum film thickness

HcminMinimum film thickness on the central line

of the contact in the x-direction

K0,K0

0,B,R0 Tait-Doolittle model constants

Gc,nc Carreau equation constants

l1,l2 Low-shear and high-shear limiting

viscosities, respectively

bK,ec,aV Tait-Doolittle model constants

DT Temperature variation = T-T0

Dimensionless parameters

M, L Dimensionless Moes–Venner parameters

P ¼ pph

q ¼ qq0

l ¼ ll0

H ¼ hRa2

X ¼ xa Y ¼ y

a Z ¼za :Solids p and s

zh :Lubricant

1 Introduction

Lubrication of machine elements plays an important role in

the proper functioning of a mechanical system, preventing

metal-to-metal contact which may damage the system and

lead to failure. This also reduces the energy consumption of a

machine since friction forces become less important when

surface separation is ensured by a complete lubricant film. In

many operating conditions, the pressure generated in the

lubricant film is high enough to induce a considerable elastic

deformation of the contacting bodies. This lubrication

regime is known as ElastoHydroDynamic (EHD).

The classical isothermal EHD theory assumes a Newto-

nian lubricant and an isothermal operating regime. But, in

most operating conditions these assumptions fail in repre-

senting the physical reality of the contact. In fact, lubricants

show a non-Newtonian response at high shear stresses.

Moreover, at high speed and/or load operating conditions, or

at high sliding velocities, the heat generation in the lubricant

film becomes important. Therefore, temperature effects

cannot be considered as negligible. In fact, an increase in

temperature leads to a decrease in the viscosity of the

lubricant which induces a reduction in both film thickness

and traction force. The interest in thermal effects for EHD

lubrication first appeared with the pioneering theoretical

work of Cheng [1, 2]. The first full numerical solution for the

point contact problem was obtained by Zhu and Wen [3].

Since then, several authors have proposed different methods

to deal with this problem assuming a Newtonian or a non-

Newtonian lubricant such as Kim and Sadeghi [4], Guo et al.

[5] and also Liu et al. [6] who solved the three-dimensional

energy equation to determine the temperature variations

throughout the lubricant film. An alternative method was

proposed by Kim et al. [7] who reduced the 3-D heat transfer

problem to a 2-D one by assuming a parabolic distribution of

the temperature across the film thickness. However, the

parabolic temperature profile simplification leads to tem-

perature predictions that are not accurate especially at the

inlet of the contact. The reason lies in the complex thermal

convective effects which are associated with important

reverse flows in this area. This is why, in this work, a full

resolution of the energy equation applied to both contacting

solids and the lubricant film is adopted.

Until now, most of the thermoelastohydrodynamic

lubrication (TEHL) numerical models used rheological

models such as the sinh-law [8] model for non-Newtonian

behaviour, the Barus [9] and Roelands [10] equations for

viscosity–pressure–temperature dependence and also the

Dowson and Higginson [11] relationship for density–

pressure–temperature dependence. Although these models

have been used for a long time in the field of Tribology, it

is recognized that they do not always describe the real

physical behaviour of a typical lubricant. This may explain

(in part) the discrepancy observed between numerical

results and experiments. The only reason for their extensive

use was their simple mathematical form, well adapted for

EHL solvers. In this paper, more realistic models are used

to define the rheology of both Newtonian and non-New-

tonian lubricants. This leads to a good agreement between

numerical results and experimental data.

2 Numerical Model

In this section, the global numerical procedure used for the

modelling of the TEHL of circular contacts is described.

The goal is to model a lubricated contact between a sphere

and a plane under a prescribed external load. Both con-

tacting bodies are elastic and have constant surface

velocities. Surface separation is ensured by a complete

lubricant film.

2.1 Thermal EHL Equations

The Reynolds equation for a steady-state point contact

lubricated with a generalized Newtonian lubricant under

thermal EHL operating regime and unidirectional surface

velocities in the x-direction is given by Yang and Wen [12]:

o

ox

qg

e

h3 op

ox

þ o

oy

qg

e

h3 op

oy

¼ 12

o

oxqUmhð Þ;

ð1Þ

42 Tribol Lett (2008) 30:41–52

123

Page 3: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

where

Um ¼ upþus

2qg

e¼ 12

geq0e

g0e q00e

q ¼ q0ege usupð Þþqeup½ Um

qe ¼ 1h

R h

0qdz

q0e ¼ 1h2

R h0

qR z

0dz0

g dz q00e ¼ 1h3

R h0

qR z

0z0dz0

g dz1ge¼ 1

h

R h

0dzg

1g0e¼ 1

h2

R h

0zdzg

Note that this equation accounts for the variations of

both density and viscosity across the film thickness as can

be seen in the integral terms. In fact, the changes in density

are due to temperature variations across the lubricant film

whereas the changes in viscosity stem from both

temperature and (when a generalized Newtonian lubricant

is considered) shear rate variations across the film.

Moreover, both density and viscosity are allowed to vary

with pressure and temperature throughout the lubricant

film. Equation 1 is solved on the 2-D contact area Xc (see

Fig. 1). Indices p and s correspond to the plane and the

sphere, respectively, and g is the generalized Newtonian

viscosity defined as:

g ¼ se

f se; lð Þ with f se; lð Þ ¼ _ce:

where se and _ce are the equivalent shear stress and shear

rate, respectively, given by:

se ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffis2

xz þ s2yz

q_ce ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi_c2

xz þ _c2yz

q

And the function f could stand for any generalized

Newtonian model, e.g. Eyring [8] or Carreau [13]. In the

case of a Newtonian lubricant, the generalized Newtonian

viscosity is replaced by the normal Newtonian viscosity,

g = l. This way, Eq. 1 becomes the classical Reynolds

[14] equation. Thus, the same model can be used for both

Newtonian and non-Newtonian lubricants.

The boundary conditions for the previous equation

consist in putting to zero the pressure on the boundary of

the contact area qXc. As for the free exit boundary prob-

lem, a Heavyside function Hs(p) is defined as:

Hs pð Þ ¼ 1 if p [ 0

0 if p 0

Let p+ = p 9 Hs(p) be the positive part of the pressure

distribution p. It is used to define both the density–pressure

and viscosity–pressure dependence and it also serves as a

normal load in the boundary condition for the elastic

deformation calculation. This is sufficient to handle the

free boundary problem at the outlet area of the contact as

shown in [15].

The film thickness equation is given by:

h x; yð Þ ¼ h0 þx2 þ y2

2R w x; yð Þ: ð2Þ

The normal elastic displacement of the contacting

bodies w(x,y) is obtained by solving the linear elasticity

equations on both the sphere and the plane. The latter are

large enough compared to the contact size to be considered

as semi-infinite structures. This is why their geometry is

defined by a large 3-D cubic structure. Let X be the interior

domain of the structure, qX its boundary and Xc the part of

the upper boundary that corresponds to the contact domain

(see Fig. 1). The linear elasticity equations consist in

finding the displacement vector U on the computational

domain X such that:

div rð Þ ¼ 0 with r ¼ Ce Uð Þ and U ¼ u; v;wf g ð3Þ

with the following boundary conditions:

U ¼ 0 at the bottom boundaryoXb

rn ¼ pþ at the contact area boundaryXc

rn ¼ 0 elsewhere

8<:

To simplify the previous model and reduce the

computational effort, an equivalent problem is defined

to replace the two identical problems of the elastic

deformation calculation for both contacting bodies. This

equivalent problem is obtained by the superposition of the

two linear problems. Let (Ep,tp) and (Es,ts) be the material

properties (Young’s modulus and Poisson’s ratio) of the

two bodies. The equivalent problem is defined by applying

Eq. 3 to a body that has the following material properties:

Eeq ¼E1E2

E1 þ E2

teq ¼tpEs þ tsEp

E1 þ E2

:

For more details about this model the reader is referred to

[15].

X

ZY

Contactarea

bΩ∂

Ω

U=0

Fig. 1 Three-dimensional domain for the linear elasticity problem

Tribol Lett (2008) 30:41–52 43

123

Page 4: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

The energy equations for the two solids p and s are given

by:

cpqpupoTox ¼ kp

o2Tox2 þ o2T

oy2 þ o2Toz2

csqsus

oTox ¼ ks

o2Tox2 þ o2T

oy2 þ o2Toz2

8<: ð4Þ

The geometrical domains of solids pand sare taken as

infinite layers with a finite thickness sufficiently large to

have a zero temperature gradient in the z-direction at the

depth d. The origin of the global coordinates system is

located at the centre of the Hertzian contact area on the

plane’s surface. The boundary conditions for the two solids

are then given by:

T xin; y; zð Þ ¼ T0; T x; y;dð Þ ¼ T0;T x; y; hþ dð Þ ¼ T0:

As for the lubricant film, neglecting the heat conduction

in the x- and y-directions and convection in the z-direction,

the energy equation is given by:

where the velocity components uf and vf are given,

according to [12], by:

uf ¼ up þ opox

R z

0z0dz0

g ge

g0ehR z

0dz0

g

h iþ ge usupð Þ

h

R z

0dz0

g

vf ¼ opoy

R z

0z0dz0

g ge

g0ehR z

0dz0

g

h i :

One should note that there are two energy sources in this

equation: compressive heating/cooling and shear heating.

In general, the shear source is more important than the

compressive one. Equation 5 requires a boundary condition

in the inlet where uf C 0 while for negative values of uf the

boundary condition is unnecessary:

T xin; y; zð Þ ¼ T0 if uf xin; y; zð Þ 0:

On the two lubricant–solid interfaces, heat flux

continuity boundary conditions must be satisfied:

k oToz

z¼0þ¼ kp

oToz

z¼0

k oToz

z¼h¼ ks

oToz

z¼hþ

(

Finally, the load balance equation is given by:ZXc

pþdX ¼ L: ð6Þ

Equation 6 is used to ensure the correct load is applied.

One shiuld note that Eq. 1 is 2-D whereas Eqs. 3–5 are

three dimensional. The domains of application of these

equations are given in the following section.

2.2 Numerical Procedure

In practice, all the above equations are solved in their non-

dimensional form. The symmetry of the problem is taken

into account, reducing thus its size to the half. The space

dimensions are non-dimensonalized with respect to the

Hertzian contact radius except for the lubricant film where

the height z is non-dimensonalized with respect to the film

thickness h. Therefore, the geometrical domain of the

lubricant film extends like both the contact area Xc and the

thermal domains for solids p and s from -4.5 B X B 1.5

and -3 B Y B 0, while its height Z [ [0,1]. The height of

solids p and s is D = d/a. The semi-infinite structure for

the linear elasticity problem should be large enough com-

pared to the contact area. This stems from the physical

reality of an EHD contact where the size of the contact is

very small compared to the size of the contacting bodies. A

size of 60 9 30 9 60 was shown to be suitable in [15].

Unstructured variable tetrahedral meshing is used for both

the EHD and thermal problems’ geometries. This allows

the use of fine meshing only where needed, leading thus to

small size systems. In fact, the EHD problem consists of a

total number of 40,000 degrees of freedom whereas for the

thermal problem the total number raises to 65,000.

Lagrange second-order elements are used for both the

elastic and thermal problems (linear elasticity and energy

equations) whereas fifth-order elements are used for the

hydrodynamic problem (Reynolds equation). For more

details about the EHD model, the reader is referred to [15].

A classical finite-element procedure is applied to Eqs. 1,

3–5. For the energy equations, a Streamline Upwind Petrov

Galerkin formulation [16] is used to stabilize the solution

for convection-dominated problems. In fact, whenever the

Peclet number Pe exceeds 1, oscillatory behaviour of the

solution is likely to occur. When convection becomes

dominant, Galerkin discretization is no longer appropriate

and gives rise to spurious oscillations in the solution. One

qc ufoT

oxþ vf

oT

oy

¼ k

o2T

oz2 T

qoqoT

uf

op

oxþ vf

op

oy

|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflzfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflCompressive heating=cooling

þ gouf

oz

2

þ ovf

oz

2" #

|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflzfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflShear heating

; ð5Þ

44 Tribol Lett (2008) 30:41–52

123

Page 5: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

way to get rid of these oscillations is by using streamline

upwind discretizations (also known as ‘‘artificial diffusion’’

techniques). Starting with an initial value of h0 and an

initial pressure, elastic deformation, temperature and shear

stress profiles, a full-system approach is applied to Eqs. 1

and 3. In other words, they are solved simultaneously using

a Newton-Raphson resolution procedure. More details of

this method are given in [15]. Then, the three energy

equations are also solved simultaneously. An iterative

process between the pressure and temperature profiles is

repeated until both are converged, in other words, until the

maximum relative difference in either solutions between

two consecutive iterations reaches 10-3. Finally, the load

balance condition is checked. If it is verified, then the

obtained solution is the good one; otherwise, the value of

h0 is updated and the same process is repeated until the

load balance is satisfied (i.e. until the calculated load

reaches the applied one within a relative error of 10-4).

3 Results

In this section, a series of test cases under both isothermal

and thermal conditions were carried out to reveal the

importance of thermal effects on both Newtonian and non-

Newtonian lubricated contacts. The Newtonian case is

briefly discussed whereas more attention is given to the

non-Newtonian case where several results are developed

and analysed. The results are compared with experimental

data. Isothermal results are obtained using the model pre-

sented in [15].

3.1 Newtonian Lubricant

First, we shall deal with the case of a liquid with a high

Newtonian limit, owing to a low (92 kg/kmol) molecular

weight:Glycerol. This liquid is not expected to shear-thin in

the inlet region of the contact [17]. The rheological

properties and operating conditions for this fluid are given

in Table 1. The rheological properties have been derived

from refs. [18, 19] whereas the thermal properties can be

found in ref. [20]. Both viscosity and density of Glycerol

have relatively low pressure dependence. Therefore, the

simple Cheng equation [1] appears to be appropriate to

define the viscosity–pressure–temperature relationship:

l ¼ l0 exp apþ b1

T 1

T0

: ð7Þ

As for the density–pressure–temperature dependence,

the Tait equation of state [21, 22] is used. It is written for

the volume relative to the volume at ambient pressure:

V

V0

¼ 1 1

1þ K 00ln 1þ p

K0

1þ K 00

: ð8Þ

This equation provides volume variation data. The density

data is obtained by simply inverting it. The initial bulk

modulus K0 and the initial pressure rate of change of bulk

modulus K00 are assumed to vary with temperature

according to:

K0 ¼ K0R exp bKTð Þ; ð9Þ

K 00 ¼ K 00R expðb0KTÞ: ð10ÞThe volume at ambient pressure relative to the ambient

pressure volume at the reference temperature TR is

assumed to depend on temperature according to:

V0

VR

¼ 1þ aV T TRð Þ: ð11Þ

Both isothermal and thermal results were obtained for

pure rolling and rolling-sliding conditions for a contact

between a steel ball and a glass plane. For the pure rolling

case the mean entrainment velocity covers the range of 0.3

to 4.75 m/s while for the rolling-sliding conditions it keeps

a constant value of 0.38 m/s with a slide-to-roll ratio (SRR)

varying from 0 to 1.8. This case was deliberately chosen to

correspond to a lightly loaded contact (ph = 0.5 GPa) to

Table 1 Lubricant properties

and operating conditions for the

Newtonian test cases

Lubricant properties Material properties Operating conditions

l0 = 0.2803 Pa s qp = 2510 kg/m3 qs = 7850 kg/m3 T0 = TR = 313 K

a = 5.4 GPa-1 kp = 1.114 W/m K ks = 46 W/m K R = 12.7 mm

b = 7468.75 K cp = 858 J/Kg K cs = 470 J/kg K L = 30 N

q0 = 1260 kg/m3 Ep = 81 GPa Es = 210 GPa ph = 0.5 GPa

k = 0.29 W/m K tp = 0.208 ts = 0.3

c = 2400 J/kg K

K0R = 12.43 GPa

bK = 0.0035 K-1

K00R = 4.5432

b0K = 0.0018 K-1

av = 5.2 9 10-4 C-1

Tribol Lett (2008) 30:41–52 45

123

Page 6: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

show that thermal effects are not restricted to highly loaded

contacts and high speeds and sliding velocities. These can

be observed even for lightly loaded contacts with moderate

speed conditions.

Figure 2 shows the central and minimum film thickness

curves as a function of the mean entrainment velocity

under pure rolling conditions. One should note that, in this

case, isothermal and thermal results are almost the same up

to 1 m/s. Beyond this speed, the two solutions diverge from

each other due to thermal effects which become important

at high speeds, even under pure rolling regime. Also note

the good agreement between the numerical results and

experimental data. At low speed (here, less than 1 m/s),

both isothermal and thermal results show a very good

agreement with the experimental results, while at high

speed, the isothermal results show some discrepancy with

experiments. On the other hand, thermal results are in good

agreement with experimental ones under any operating

conditions. Finally, note the change in the slope of the

thermal film thickness curves beyond the speed limit of

1 m/s. This is characteristic of the appearance of thermal

thinning.

Figure 3 shows the central and minimum film thick-

ness curves for a constant mean entrainment velocity

(Um = 0.38 m/s) as a function of the SRR. An isothermal

approach predicts a constant film thickness with respect

to the SRR. This is to be expected since the classical

Reynolds equation depends only on the mean entrainment

velocity. On the other hand, a thermal approach shows a

clear decrease in the film thickness when the SRR increa-

ses. This is because when the sliding velocity becomes

important, shear heating acts to reduce the viscosity of the

lubricant which leads to a decrease in the film thickness.

This is observed in both thermal and experimental results

which exhibit a better agreement compared to isothermal

and experimental ones.

3.2 Non-Newtonian Lubricant

For the non-Newtonian case, a lubricant which shear-thins

for the inlet conditions is used. It is formed from a

mixture of Squalane and 15 wt.% of PolyIsoPrene (PIP).

The rheology of this lubricant is much more complex than

Glycerol and requires more advanced rheological models

for an accurate determination of the changes in viscosity

and density with respect to the variations in pressure,

temperature and shear stress. The rheological properties

and operating conditions for the test cases of this lubri-

cant are given in Table 2. The rheological properties are

taken from ref. [17] whereas the thermal properties can be

found in ref. [20]. Again, the Tait equation of state is

used for density–pressure–temperature dependence. As for

the viscosity–pressure–temperature dependence, the free

volume model is used. The Doolittle equation [23] is

given by:

l ¼ lR exp BR0

V1V1R

VVR R0

V1V1R

1

1 R0

!" #; ð12Þ

where the viscosity at the reference state is lR = l(TR,

p = 0) and the relative occupied volume with respect to the

reference state is:

V1V1R

¼ 1þ e0 T TRð Þ: ð13Þ

Finally, the Carreau [13] model is used to represent the

shear-thinning behaviour of Squalane + PIP. This model

takes into account the second Newtonian plateau that

Fig. 2 Film thickness as a function of the mean entrainment velocity

for the Glycerol case under pure rolling conditions (ph = 0.5 GPa)

Fig. 3 Film thickness as a function of the SRR for the Glycerol case

with a constant mean entrainment velocity (Um = 0.38 m/s, ph =

0.5 GPa, M = 36.8 and L = 2.3)

46 Tribol Lett (2008) 30:41–52

123

Page 7: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

occurs at very high shear rates. A modified version

provided by Bair [24] is given under the following form:

g ¼ l2 þl1 l2

1þ se

Gc

bc

1nc1

bc

: ð14Þ

where: bc ¼ exp 0:657 0:585 ln ncð Þð Þ:Equation 14 is a good approximation for the classical

Carreau law for values of n ranging from 0.3 to 0.75, which

is the range of interest in EHL applications. l1 and l2 are

allowed to vary with temperature and pressure according to

Eqs. 8–13. The high-shear limiting viscosity at ambient

pressure l2,0 is considered to be the ambient pressure

viscosity of pure Squalane and the shear-thinning of

Squalane is neglected. This could result in a viscosity

function that is slightly less shear dependent than the

measurements would predict.

Both isothermal and thermal results were obtained for

pure rolling and rolling-sliding conditions for a contact

between a steel ball and a glass plane. For the pure rolling

case the mean entrainment velocity covers the range of 1.0

to 4.65 m/s while for the rolling-sliding conditions it keeps

a constant value of 1.47 m/s with a varying SRR. Both

thermal and non-Newtonian effects on pressure, film

thickness and traction are investigated.

3.2.1 Temperature

Before discussing pressure, film thickness and traction

results, an investigation of temperature variations in the

lubricant film is considered. These are, in part, responsible

for the reduction in viscosity that leads to a decrease in

both the film thickness and the traction coefficients.

Figure 4 shows the temperature variation (DT) profiles

Table 2 Lubricant properties

and operating conditions for the

non-Newtonian test cases

Lubricant properties Material properties Operating conditions

l1,0 = l1,R = 0.0705 Pa.s av = 7.52 9 10-4 C-1 qp = 2,510 kg/m3 TR = T0 = 313 K

l2,0 = l2,R = 0.0157 Pa.s K00R = 11.29 kp = 1.114 W/m K R = 12.7 mm

Gc = 0.01 MPa b0K = 0 K-1 cp = 858 J/kg K L = 23 N

nc = 0.8 K0R = 8.375 GPa Ep = 81 GPa ph = 0.47 GPa

B = 4.2 bK = 0.006765 K-1 tp = 0.208

R0 = 0.658 q0 = 818 kg/m3 qs = 7,850 kg/m3

e0 = -9.599 9 10-4C-1 k = 0.13 W/m K ks = 46 W/m K

c = 1,700 J/kg K cs = 470 J/kg K

Es = 210 GPa

ts = 0.3

Fig. 4 Temperature profiles across the film thickness at different X locations on the central line in the x-direction for two different values of the

SRR (0.45 and 1) with Squalane + PIP used as lubricant (Um = 1.47 m/s, ph = 0.47 GPa, M = 30 and L = 7.8)

Tribol Lett (2008) 30:41–52 47

123

Page 8: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

across the film thickness at different X locations on the

central line in the x-direction for two different values of the

SRR (0.45 and 1).

The temperature variation is more important in the case

of SRR = 1 revealing thus the higher heat generation due to

shear heating. We can also see that in the two cases, the

temperature on the plane’s surface is higher than on the

sphere’s surface. This is to be expected since the plane is

made out of glass which has much lower thermal diffusivity

(k/qc) and effusivity (ffiffiffiffiffiffiffiffikqcp

Þ than steel. So, if the two sur-

faces had the same velocity the ball’s surface is expected to

have a lower temperature because steel has a higher ability

to exchange energy with its surrounding than glass. And

since for positive values of the SRR, the surface velocity of

the ball is higher than that of the plane, and knowing that

steel has a higher volumetric heat capacity (qc) than glass,

the heat removed from the ball by convection is also more

important than for the plane. This makes the difference in

surface temperature between the two bodies even more

pronounced. Finally, note the increase in the temperature of

the lubricant as it enters the contact until it reaches its

maximum in the central area before decreasing as the

lubricant goes out of the contact. In this outlet region, we

can also see a reverse in the orientation of the temperature

variation parabola which reveals the importance of the

compressive cooling effect that occurs in this area where the

pressure gradient is negative.

Now, let us investigate the effects of these variations in

temperature along with the generalized Newtonian effect

on pressure, film thickness and traction results.

3.2.2 Pressure

First, in order to isolate the effect of the non-Newtonian

behaviour of the lubricant on the pressure profile, let us

have a look on the plot of the isothermal pressure distri-

bution along the central line in the x-direction for a

constant mean entrainment velocity of 1.47 m/s and dif-

ferent SRRs. Figure 5 shows that, globally, the pressure

distribution is not affected by the increase of sliding. The

only noticeable difference can be observed in the pressure

spike’s region. The latter loses height when the SRR is

increased. This was already observed in a previous work of

the authors [25].

The same numerical cases were run under thermal

conditions. This reveals the combined effect of temperature

and non-Newtonian behaviour of the lubricant on pressure.

And mostly, by comparison with the corresponding previ-

ous isothermal results, it allows to isolate the temperature

effects.

Figure 6 shows that when thermal effects are taken into

account, globally, the pressure exhibits a decrease in the

central contact area. Since the load balance should be

satisfied, this is generally compensated by an increase of

the pressure in the width of the contact. This was observed

in experiments by Jubault et al. [26]. It is also shown in

Fig. 7 where we can note a pressure increase in the contact

width in the y-direction with the increase of the SRR.

Concerning the pressure spike, not only does it lose

height due to shear-thinning when the SRR is increased,

but, due to thermal effects, it also gains width and moves

towards the centre of the contact. This has a direct con-

sequence on the shape of the film thickness profile

especially in the horseshoe area at the outlet of the contact.

The variations in the film thickness shape are discussed

more in detail in the following section.

3.2.3 Film Thickness

For pure rolling tests, the central and minimum film

thickness curves for isothermal, thermal and experimental

results are shown in Fig. 8. As for the Newtonian case, up

to a given velocity limit (in this case 2 m/s), isothermal and

thermal results are practically the same. Beyond this value,

the two curves diverge from each other revealing thus the

appearance of important shear heating. Finally, note the

exceptional agreement between thermal and experimental

results.

Fig. 5 Non-Newtonian effects

on pressure for the

Squalane + PIP case (Um

= 1.47 m/s, ph = 0.47 GPa,

M = 30 and L = 7.8)

48 Tribol Lett (2008) 30:41–52

123

Page 9: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

Figure 9 shows the film thickness curves as a function of

the SRR. In addition to the central and minimum film

thicknesses, the minimum film thickness on the central line

in the x-direction curves for isothermal, thermal and

experimental approaches are shown. The film thickness

continuously decreases with the increase of the SRR for

both isothermal and thermal approaches. This decrease is

more important in the thermal case since the combined

non-Newtonian and thermal effects are superposed, and

both have a thinning effect on the film thickness. Also note,

especially in the case of central film thicknesses, the good

agreement between thermal and experimental results.

In both the thermal and experimental cases, when the

SRR is increased, the minimum film thickness on the

central line in the x-direction hcminapproaches the global

minimum film thickness hmin. Therefore, the horseshoe

shape at the outlet of the contact, which originally has large

ends and a narrow central region, gains in width on its

central part and starts having an almost constant width. The

change in shape of the horseshoe due to thermal effects was

also observed in the experiments of Jubault et al. [26]. By

examining Fig. 10, one can see in the contour plots for the

thermal cases that when the SRR grows, the horseshoe

loses width on its end parts and becomes larger on its

central part. Thus, it comes closer to having a global

constant width. This could not be observed by a simple

isothermal approach where the difference between hcminand

hmin is almost constant whatever the SRR was (see Fig. 9).

This is also observed on the isothermal contour plots in

Fig. 11 where the horseshoe has practically the same shape

for the different SRRs.

3.2.4 Traction

Figure 12 shows the isothermal, thermal and experimental

traction curves as a function of the SRR for two different

constant mean entrainment velocities of 0.74 and 1.47 m/s.

It is clear that an isothermal approach is not appropriate

for estimating friction in an EHL contact. In fact, it

Fig. 7 Combined thermal and non-Newtonian effects on the width

of the contact for the Squalane + PIP case (Um = 1.47 m/s, ph =

0.47 GPa, M = 30 and L = 7.8)

Fig. 8 Film thickness curves as a function of the mean entrainment

velocity under pure rolling conditions for the Squalane + PIP case

(ph = 0.47 GPa)

Fig. 6 Combined non-

Newtonian and thermal effects

on pressure for the

Squalane + PIP case (Um

= 1.47 m/s, ph = 0.47 GPa,

M = 30 and L = 7.8)

Tribol Lett (2008) 30:41–52 49

123

Page 10: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

overpredicts friction coefficients whereas a thermal

approach shows a good agreement with experimental

data.

Finally, one should note that the isothermal curves

predict a higher traction coefficient for the higher mean

entrainment velocity whatever the SRR was. This does not

reflect the physical reality since the experimental points

(and the thermal curves) show that this is true only up to a

certain value of the SRR (here &0.5). Beyond this value,

the tendency is inverted and the friction coefficient for the

higher mean entrainment velocity becomes lower. This

reveals the appearance of important thermal effects at high

speed operating conditions. In this case, beyond SRR &0.5 and for Um = 1.47 m/s, the latter have more influence

on the traction coefficient than frictional shear, whereas for

Um = 0.74 m/s, frictional shear is dominant.

Fig. 9 Film thickness curves as

a function of the SRR for a

constant mean entrainment

velocity for the Squalane + PIP

case (Um = 1.47 m/s,

ph = 0.47 GPa, M = 30 and

L = 7.8)

Fig. 10 Thermal film thickness

contour plots as a function of

the SRR for the Squalane + PIP

case (Um = 1.47 m/s,

ph = 0.47 GPa, M = 30 and

L = 7.8)

Fig. 11 Isothermal film

thickness contour plots as a

function of the SRR for the

Squalane + PIP case (Um

= 1.47 m/s, ph = 0.47 GPa,

M = 30 and L = 7.8)

50 Tribol Lett (2008) 30:41–52

123

Page 11: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

4 Conclusion

This paper presents a finite element model for TEHL of

circular contacts lubricated with either Newtonian or shear-

thinning lubricants. Some typical test cases were run under

isothermal and thermal operating regimes for both pure

rolling and rolling-sliding conditions. It is shown that

beyond a certain limit of speed operating conditions, it is

necessary to take into account shear heating for a good

estimation of film thickness and, more importantly, friction

coefficients. Shear-thinning and thermal effects are shown

to have a film-thinning effect when the sliding velocity is

increased. They also modify both pressure and film thick-

ness distributions. In fact, shear-thinning effects tend to

decrease the pressure spike’s height without any significant

change in the shape of the film thickness distribution when

the SRR is increased. As for thermal effects, not only do

they act to decrease the height of the pressure spike, but

they also provide a gain in its width and its approach

towards the centre of the contact. This leads to a change in

the film thickness profile, especially in the outlet region

where the horseshoe shape becomes wider at its central part

and less large at its ends. Thus, it has an almost constant

width compared to the isothermal shape where the horse-

shoe is large at the ends and narrow in the central region.

All the results were compared to experimental data show-

ing a better agreement between thermal results and

experiments especially at high speeds or sliding velocities.

A remarkable agreement was obtained for friction results

where the isothermal approach was shown to overestimate

the friction coefficients in the contact.

Finally, the thermal conductivity of organic liquids is

known to increase with pressure. In this work, the ambient

pressure value has been used, neglecting the pressure

dependence. This does not significantly affect the results

since only moderately loaded contacts are treated here. But,

for future works, when dealing with highly loaded contacts,

the influence could become more important and a more

complete analysis will be carried out. This is expected to

reduce the temperature increase throughout the film

thickness.

Acknowledgements The authors thank Prof. E. Ioannides (SKF

Group Technical Director) for his kind permission to publish this

work. They also wish to express their gratitude to the French Ministry

of National Education and Scientific Research for partially financing

this study.

References

1. Cheng, H.S., Sternlicht, B.: A numerical solution for the pressure,

temperature and film thickness between two infinitely long,

lubricated rolling and sliding cylinders, under heavy loads.

ASME J. Basic Eng. 87, 695–707 (1965)

2. Cheng, H.S.: A refined solution to the thermal-elastohydrody-

namic lubrication of rolling and sliding cylinders. ASLE Trans 8,

397–410 (1965)

3. Zhu, D., Wen, S.: A full numerical solution for the thermo-el-

astohydrodynamic problem in elliptical contacts. ASME J.

Tribol. 106, 246–254 (1984)

4. Kim, K.H., Sadeghi, F.: Three-dimensional temperature distri-

bution inEHD lubrication: Part I—circular contact. ASME J.

Tribol. 114, 32–41 (1992)

5. Guo, F., Yang, P., Qu, S.: On the theory of thermal elastohy-

drodynamic lubrication at high slide-roll ratios—circular glass-

steel contact solution at opposite sliding. ASME J. Tribol. 123,

816–821 (2001)

6. Liu, X., Jiang, M., Yang, P., Kaneta, M.: Non-Newtonian thermal

analyses of point EHL contacts using the eyring model. ASME J.

Tribol. 127, 70–81 (2005)

7. Kim, H.J., Ehret, P., Dowson, D., Taylor, C.M.: Thermal elas-

tohydrodynamic analysis of circular contacts, Part 1: Newtonian

model. IMechE J. Eng. Tribol. 215(J), 339–352 (2001)

8. Eyring, H.: Viscosity, plasticity and diffusion as examples of

absolute reaction rates. J. Chem. Phys. 4, 283–291 (1936)

9. Barus, C.: Isothermals, isopiestics and isometrics relative to

viscosity. Am. J. Sci. 45, 87–96 (1893)

10. Roelands, C.J.A.: Correlational Aspects of the viscosity-temper-

ature-pressure relationship of lubricating oils. PhD thesis,

Technische Hogeschool Delft, The Netherlands (1966)

11. Dowson, D., Higginson, G.R.: Elastohydrodynamic Lubrication.

The Fundamental of Roller and Gear Lubrication. Pergamon,

Oxford (1966)

12. Yang, P., Wen, S.: A generalized Reynolds equation for non-

Newtonian thermal elastohydrodynamic lubrication. ASME J.

Tribol. 112, 631–636 (1990)

13. Carreau, P.J.: Rheological equations from molecular network

theories. Trans. Soc. Rheol. 16(1), 99–127 (1972)

14. Reynolds, O.: On the theory of the lubrication and its application

to Mr Beauchamp Tower’s experiments, including an experi-

mental determination of the viscosity of olive oil. Philos. Trans.

R. Soc. 177, 157–234 (1886)

Fig. 12 Traction curves as a function of the SRR for two different

mean entrainment velocities of 0.74 m/s (M = 50 and L = 6.5) and

1.47 m/s (M = 30 and L = 7.8) for the Squalane + PIP case

(ph = 0.47 GPa)

Tribol Lett (2008) 30:41–52 51

123

Page 12: Thermal Elastohydrodynamic Lubrication of Point Contacts Using a Newtonian Generalized Newtonian Lubricant

15. Habchi, W., Eyheramendy, D., Vergne, P., Morales-Espejel, G.:

A full-system approach of the elastohydrodynamic line/point

contact problem. ASME J. Tribol. 130(2), (2008, to appear)

16. Brooks, A.N., Hughes, T.J.R.: Streamline-upwind/Petrov-Galer-

kin formulations for convective dominated flows with particular

emphasis on the incompressible Navier-Stokes equations. Comp.

Methods Appl. Mech. Eng. 32, 199–259 (1982)

17. Bair, S.: High-Pressure Rheology for Quantitative Elastohydro-

dynamics. Elsevier Science, Amsterdam (2007)

18. Cook, R.L., King, H.E., Herbst, C.A., Herschback, D.R.: Pressure

and temperature dependent viscosity of two glass forming liquids:

glycerol and dibutyl phthalate. J. Chem. Phys. 100(7), 5178–5189

(1994)

19. Bridgman, P.W.: The Physics of High Pressure. Dover, New

York (1970)

20. Lide, D.R.: CRC Handbook of Chemistry and Physics, 85th edn.

CRC press, Boca Raton-Florida (2004)

21. Hirschfelder, J.O., Curtiss, C.F., Bird, R.B.: Molecular Theory of

Gases and Liquids. Wiley, New York (1954)

22. Hogenboom, D.L., Webb, W., Dixon, J.D.: Viscosity of several

liquid hydrocarbons as a function of temperature, pressure and

free volume. J. Chem. Phys. 46(7), 2586–2598 (1967)

23. Doolittle, A.K.: Studies in Newtonian flow II, the dependence of

the viscosity of liquids on free-space. J. Appl. Phys. 22, 1471–

1475 (1951)

24. Bair, S.: A rough shear-thinning correction for EHD film thick-

ness. STLE Tribol. Trans. 47, 361–365 (2004)

25. Habchi, W., Demirci, I., Eyheramendy, D., Morales-Espejel, G.,

Vergne, P.: A finite element approach of thin film lubrication in

circular EHD contacts. Tribol. Int. 40, 1466–1473 (2007)

26. Jubault, I., Molimard, J., Lubrecht, A.A., Mansot, J.L., Vergne,

P.: In situ pressure and film thickness measurements in rolling/

sliding lubricated point contacts. STLE Tribol. Lett. 15(4), 421–

429 (2003)

52 Tribol Lett (2008) 30:41–52

123