scispace - formally typeset
Open AccessJournal ArticleDOI

Analysis of Mixed Lubrication Effects in Simulated Gear Tooth Contacts

Reads0
Chats0
TLDR
In this article, the authors used a transient analysis technique for point contact elastohydrodynamic lubrication (EHL) problems based on a formulation that couples the elastic and hydrodynamic equations.
Abstract
The paper presents results obtained using a transient analysis technique for point contact elastohydrodynamic lubrication (EHL) problems based on a formulation that couples the elastic and hydrodynamic equations. Results are presented for transverse ground surfaces in elliptical point contact that show severe film thinning and asperity contact at the transverse limits of the contact area. This thinning is caused by transverse leakage of the lubricant from the contact in the remaining deep valley features between the surfaces. A comparison is also made between the point contact results on the entrainment center line and the equivalent line contact analysis. The extent of asperity contact is shown to be dependent on the Hertzian contact aspect ratio. It is also shown that transverse waviness (superimposed on the roughness) of even relatively small amplitude can lead to large increases in asperity contact rates over all waviness peaks in the contact.

read more

Content maybe subject to copyright    Report

M. J. A. Holmes
H. P. Evans
R. W. Snidle
Cardiff School of Engineering,
Cardiff CF24 0YF, UK
Analysis of Mixed Lubrication
Effects in Simulated Gear Tooth
Contacts
The paper presents results obtained using a transient analysis technique for point contact
elastohydrodynamic lubrication (EHL) problems based on a formulation that couples the
elastic and hydrodynamic equations. Results are presented for transverse ground surfaces
in elliptical point contact that show severe film thinning and asperity contact at the
transverse limits of the contact area. This thinning is caused by transverse leakage of the
lubricant from the contact in the remaining deep valley features between the surfaces. A
comparison is also made between the point contact results on the entrainment center line
and the equivalent line contact analysis. The extent of asperity contact is shown to be
dependent on the Hertzian contact aspect ratio. It is also shown that transverse waviness
(superimposed on the roughness) of even relatively small amplitude can lead to large
increases in asperity contact rates over all waviness peaks in the contact.
DOI: 10.1115/1.1828452
Introduction
The lubrication mechanism primarily responsible for the pro-
tection of gear tooth surfaces from wear and surface distress is
elastohydrodynamic lubrication EHL. In the case of very smooth
surfaces such as those found in rolling element bearings, for ex-
ample the EHL mechanism can generate oil films which are thick
compared to the height of roughness features present on the sur-
faces. Under these conditions the thickness of the oil film may be
calculated, with reasonable accuracy, using the well-known for-
mula of Dowson and Higginson 1. An important feature of gear
tooth contacts, however, is that the surfaces produced by present
day manufacturing methods have roughness features that are sig-
nificantly greater than the oil film predicted by this formula. Con-
sequently gears tend to operate in a regime described as ‘mixed’
or ‘micro’ EHL in which there is a significant interaction of
roughness asperities on the two surfaces. In theoretical solutions
of both the dry and micro EHL situations the presence of rough-
ness leads to significant rippling of the contact pressure distribu-
tion with maximum values far in excess of the Hertzian values
expected when the surfaces are perfectly smooth. Micro EHL so-
lutions also indicate the presence of very thin films at asperity
encounters within the overall rolling/sliding contact. Two practical
problems associated with roughness effects and film thinning in
gears are micropitting rolling contact fatigue on the scale of sur-
face asperities and scuffing scoring which is related to the fail-
ure of the elastohydrodynamic system. In order to gain a much
clearer understanding of these failure mechanisms it is necessary
to develop a full theoretical model of lubrication of gear contacts
under rough surface/thin film conditions. Such a model must take
account of the real operating conditions of gears in terms of loads,
speeds, surface roughness and lubricant properties, and be able to
predict pressures, local film thickness, temperatures and friction
between the teeth. A further important feature that must be con-
sidered when roughness is present is the time-dependent effect of
roughness: this occurs when roughness features move relative to
the overall contact.
The paper presents results from the numerical modeling of tran-
sient rough surface point contact problems obtained using a new
coupled numerical formulation for solving the elastohydrody-
namic EHL point contact problem described and validated else-
where 2,3. The paper focuses attention on the strong side leak-
age effects that take place at the edges of contacts that have a
surface finish transverse to the direction of rolling/sliding such as
that in conventional involute gears. The configuration chosen for
analysis is that of a gear simulation disk rig which gives rise to an
elliptical contact. In the smooth surface case the EHL contact
adopts a self-sealing configuration by developing side constric-
tions in the form of the familiar horseshoe shape seen in optical
interferometry experiments. When transverse roughness features
are present, however, this mechanism is unable to seal the
pressure-driven transverse flow in the valley features because the
closest that the surfaces can be brought together is determined by
the physical contact of asperity tip features. Even in this extreme
configuration the composite valley features on the surfaces remain
open and unsealed, and lubricant can easily escape from the con-
tact area in the transverse direction along these valleys. When oil
is lost from the contact due to this sideways leakage mechanism
the entrainment of lubricant under the downstream micro contacts
is progressively weakened at each successive following contact.
This model for EHL failure was proposed earlier by the authors
4 in response to their experimental observation that initial scuff-
ing failure invariably occurs at the transverse edges of such con-
tacts 5. The detailed results of micro-EHL analysis presented
here add further evidence in support of the model. It is of interest
to note that the transverse edges, where failure appears to origi-
nate in the scuffing experiments, are not subject to extreme tem-
perature or extreme pressure behavior. The identification of con-
tact edges as the location of initial scuffing failure is thus a
significant observation indicating that failure of the physical
mechanism of EHL is a primary underlying cause of scuffing
in gear tooth contacts. In real gears having a finite facewidth
effective ‘contact edges’ which behave in the way suggested
are not limited to the actual face edges of the gears because of
the inevitable ‘waviness’ present on the surfaces. Results pre-
sented in the paper suggest that such waviness, of even small
amplitude, can lead to a significantly increased occurrence of film
breakdown.
Contact Analysis
The EHL problem is specified by the elastic deflection equation
written in differential form 2,
Manuscript received February 23, 2004; revision received August 16, 2004. Re-
view conducted by: M. Lovell.
Copyright © 2005 by ASMEJournal of Tribology JANUARY 2005, Vol. 127 Õ 61
Downloaded 04 Apr 2012 to 131.251.133.27. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

2
h
x
i
,y
j
x
2
2
h
x
i
,y
j
y
2
2
x
i
,y
j
x
2
2
x
i
,y
j
y
2
1
R
x
1
R
y
2
E
all k,l
f
k i,l j
p
k,l
(1)
and the non-Newtonian Reynolds equation
x
x
p
x
y
y
p
y
U
¯
h
x
h
t
0 (2)
The nonlinear dependence of viscosity and density on pressure are
taken to be given by the well known isothermal Roelands, and
Dowson and Higginson relationships, respectively,
0
exp
ln
0
/
共共
1
p
Z
1
(3)
0
1
p
1 p
(4)
For non-Newtonian situations
x
and
y
(
x
) are determined
from the lubricant’s pressure, pressure gradients, film thickness,
and surface velocities as discussed in 6. Equation 2 is dis-
cretized using linear quadrilateral finite elements with an implicit
CrankNicolson time formulation, and Eq. 1 using finite dif-
ferences for the Laplacian with the pressure coefficients, f
i,j
,
given by the analysis in 7. The equations are thus written as
k 0
n
c
A
k
p
k
k 0
n
c
B
k
h
k
R
i,j
(5)
k 0
n
c
C
k
p
k
k 0
n
c
D
k
h
k
E
i,j
(6)
where suffix k represents the nodes contributing to the assembled
equation at node (i,j) and k 0 denotes that node. A
k
and B
k
are
the pressure and film variable coefficients for the Reynolds equa-
tion 2, and n
c
is the number of neighboring nodes involved in
the formulation. Similarly C
k
and D
k
are the pressure and film
variable coefficients for the differential deflection equation 1.
Expression R
i,j
contains information from the previous timestep,
and E
i,j
contains all the contributions to the pressure summation
of Eq. 1 that are not explicitly contained on the left hand side of
Eq. 6.
Equations 5 and 6 are expressed as a pair of simultaneous
equations in the variables p
0
and h
0
A
0
p
0
B
0
h
0
R
ˆ
i,j
R
i,j
k 1
n
c
A
k
p
k
k 1
n
c
B
k
h
k
C
0
p
0
D
0
h
0
E
ˆ
i,j
E
i,j
k 1
n
c
C
k
p
k
k 1
n
c
D
k
h
k
which are solved to give a coupled iterative scheme to update the
values of all the unknown nodal values of p and h in turn, i.e.,
p
i,j
new
R
ˆ
i,j
D
0
E
ˆ
i,j
B
0
A
0
D
0
B
0
C
0
, h
i,j
new
E
ˆ
i,j
A
0
R
ˆ
i,j
C
0
A
0
D
0
B
0
C
0
(7)
This method is found to be both effective in obtaining rapid con-
vergence in low situations with rough surfaces, and extremely
robust. The rapidity with which the influence coefficients f
i,j
in
Eq. 1 fall, as the indices i and j increase from zero 7,isakey
advantage of this differential formulation of the deflection equa-
tion. This property allows the recalculation of pressure contribu-
tions to E
i,j
to be limited to those that are close to the n
c
points
used in the iteration sweep 2,8.
During rough surface transient analyses fluid film breakdown
can occur resulting in contact between the micro asperities. Where
contact occurs between the two surfaces the hydrodynamic film
thickness is zero, although in practice there will typically be a
boundary film that controls the local friction coefficient. Equation
2 arises from mass flow continuity of the fluid film, and at lo-
cations where the film thickness is zero there is no such mass
flow. The physical principle on which Eq. 2 is founded is thus
not applicable at micro contact locations. Equation 1, however,
is always applicable as it relates the pressure acting on the sur-
faces with their deflection irrespective of whether the pressure
arises from a hydrodynamic film or from direct contact of the
surfaces.
Contact situations in the iterative scheme are dealt with as fol-
lows. If the iterating equations 7 result in a negative value for
h
i,j
new
, its value is set to zero and Eqs. 7 are thus replaced by
h
i,j
new
0, p
i,j
new
E
ˆ
i,j
C
0
(8)
This effectively replaces the Reynolds equation with the boundary
condition h 0, and applies the deflection equation subject to that
boundary condition.
Thus Eqs. 7 are used at each mesh point during each iterative
sweep and are replaced by Eqs. 8 only at mesh points where the
current evaluation of Eqs. 7 yield a negative value for h
i,j
new
. The
ease with which contact conditions can be incorporated using this
approach is a further advantage of the coupled differential deflec-
tion technique.
This method can be used to solve the dry, elastic contact prob-
lem using Eq. 2 as can be seen from the results for contact
start-up analysis presented in 9. This problem involves a simul-
taneous solution of full film and dry contact areas as liquid is first
entrained into a dry contact by motion of the surfaces. The itera-
tive approach described above deals effectively with this situation,
maintaining a dry contact pressure that remains essentially Hert-
zian away from the area where the contact shape is distorted by
the entrained fluid. The comparison made in 9 with the elegant
experimental work of Glovnea and Spikes 10 for this situation
provides confirmation of the validity of the approach adopted.
Results
Behavior at the Transverse Edge of the Contact. The re-
sults presented in the current paper are based on an isothermal
analysis of elliptical contacts finished in a transverse direction. In
each case the contact is between two ellipsoidal bodies whose
surface finish is given by one of the three experimental profiles
shown in Fig. 1 solid metal below the profile. Trace A is a
Fig. 1 Profiles adopted for the surfaces used in the numerical
investigation. Profiles are offset for clarity and oriented with
metal below the curves.
62 Õ Vol. 127, JANUARY 2005 Transactions of the ASME
Downloaded 04 Apr 2012 to 131.251.133.27. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

profile taken from a well run-in transverse ground disk used in
scuffing experiments by Patching 5. Traces B and C are taken
from micropitting tests on gears and have been run for several
load stages and as a result have become run-in to some extent, but
a close examination shows that they clearly have larger asperities
than profile A. Traces A, B and C have R
a
values of 0.32
m, 0.22
m and 0.31
m, respectively. Intermediate heights that
are required as the surfaces move through the contact are obtained
using cubic spline interpolation, which ensures slope continuity at
the measured points. A comparison of the line contact behavior of
these profiles has been undertaken previously in 11 where it was
clearly seen that Profile C was the most aggressive of the three
profiles in terms of its tendency to produce high pressure ripples
and severe film thinning.
The lubricant modeled is Mobil Jet 2, a synthetic gas turbine
lubricant used in earlier scuffing experiments 5. The operating
conditions and lubricant parameters adopted are specified in Table
1 and result in a contact whose Hertzian dimension in the trans-
verse direction is four times that in the rolling/sliding direction.
The computing mesh covers the area 2.5a x 1.5a; 2b
y 2b, with mesh spacing x a/200; y b/50. The
timestep adopted was t x/2u
max
so that the faster moving
surface moves through one mesh spacing over two timesteps. The
transient analysis is started from the smooth steady state result,
shown in Fig. 2 which illustrates the pressure and film thickness
contours for the operating conditions chosen for analysis. The
Moes and Bosma dimensionless groups for the contact are M
270 and L 6. These conditions could be expected to generate
an appreciable pressure spike with a Newtonian analysis, but this
is diminished into the rudimentary shoulder feature seen in non-
Newtonian circumstances; this can be discerned in the very
closely spaced monotonic pressure contours near the exit of the
Hertzian contact area. The maximum pressure developed is 1.03
GPa which is very close to the corresponding Hertzian contact
value of 1.05 GPa. The central film thickness ‘plateau’ value is
0.48
m with a minimum value on the longitudinal center line of
0.42
m, and transverse edge constrictions where a similar mini-
mum film value of 0.43
m is developed. When the steady state
solution has been established the rough surface features, which
make the problem time dependent, are fed in with the moving
surfaces from the inlet boundary position. Because of the different
speeds of the two surfaces the time taken for both surfaces in the
contact to become fully rough is that required for the slowest of
the two surfaces to move from the inlet boundary to the exit
boundary. Once this has occurred the computation is carried on for
a further 2370 timesteps, i.e. further analysis times of 0.076 ms,
0.070 ms and 0.063 ms, respectively, for the three slide/roll ratios
adopted of 0.1, 0.25 and 0.5. The distance moved by the faster
moving surface during this further calculation is thus the same in
each case, and is equal to 5.9a.
The presence of roughness on both surfaces causes a significant
variation in both pressure and film thickness within the contact
area, and Fig. 3 illustrates this effect for one particular timestep in
the analysis for two rough surfaces having profile C with a slide/
roll ratio of
0.25. This profile was shown in 11 to have the
most aggressive EHL response in rough on rough line contact of
the three profiles considered. The figure shows the pressure and
film thickness along the entrainment centreline, y 0, and also
shows the orientation of the two rough surfaces offset below for
clarity. The deviations of pressure from the smooth surface result
for this example can be seen to be significant: maximum pressures
Table 1 Operating conditions for the point and line contact
comparisons
Point
contact Line contact
R
x
0.0191 m 0.0191 m
R
y
0.151 m
w,w
962 N 527 kN/m
a 0.335 mm 0.335 mm
b 1.31 mm
p
hz
1.05 GPa 1.0 GPa
E
227 GPa
11.1 GPa
1
5.1 GPa
1
2.27 GPa
1
0
0.005 Pas 0.0048 Pas
63.2 10
6
Pas
1.68 GPa
1
U
¯
25 m/s
0.1, 0.25, 0.5
0
10 MPa
Fig. 2 PressureÕGPa upper figure and film thicknessÕ
m
contours of the smooth surface result for the conditions ana-
lyzed. Central and minimum film thicknesses are 0.48 and
0.42
m, respectively. The heavy curve indicates a Hertzian dry
contact area.
Fig. 3 Pressure heavy curve and film thickness on the en-
trainment axis,
y
Ä0, at one timestep in the analysis of contact
between two surfaces having Profile C with
Ä0.25. Also
shown are the two rough surfaces in their contact configura-
tion offset for clarity.
Journal of Tribology JANUARY 2005, Vol. 127 Õ 63
Downloaded 04 Apr 2012 to 131.251.133.27. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

of the order of 3 GPa can be seen to develop on some asperity/
asperity contacts with corresponding extremely small film thick-
ness values. As the analysis progresses contact as defined in the
previous section occurs occasionally as asperities move past each
other. The number of timesteps in which contact occurs is calcu-
lated for each mesh point during the fully rough analysis period,
and we refer to this as the contact count, denoted Q. Figure 4
shows contours of the contact count for the case where both sur-
faces have Profile C for the three different slide/roll ratios con-
sidered. To facilitate comparisons between cases having different
sliding speeds, and thus different timesteps, the values of Q ob-
tained are normalized with respect to the total analysis time. High
contact count values indicate repeated contact instances at that
particular location as the rough surfaces pass through the corre-
sponding smooth surface Hertzian contact area indicated by the
semi ellipse. It is clear from the contour values that contact occurs
predominantly at the transverse edges of the Hertzian contact area
downstream (x 0) of the contact centerline. It is worth noting
that the profiles used for the analysis are taken from experimental
test disks and that the surface finish has been modified by the
action of plastic deformation as the contact has run from its as-
manufactured surface finish to the current state. As-manufactured
surfaces are considerably rougher, and when finished by grinding
have a near Gaussian distribution of surface heights. The height
distribution of the current profiles, shown in Fig. 1, have a degree
of skewness introduced by the running-in process, and the asperity
tips are more rounded as a result. This is the surface configuration
that corresponds to contact failure, and as such its EHL behavior
is likely to be of more engineering relevance than the more com-
putationally challenging freshly manufactured finish.
In Fig. 4 we see that calculated contact is a relatively frequent
occurrence at the transverse margins of the Hertzian area. The
mesh point for which contact occurs most frequently experiences
contact in 42 of the 2370 timesteps for which the count is carried
out, i.e., about 2% of the calculation time. This is for the interme-
diate case where
0.25. The number of timesteps for which
contact occurs at one or more mesh points during the calculation
is, however, a high proportion of the total. Contact occurs for
98%, 95% and 86% of the timesteps for the three analyses as
shown in Table 2.
There is little difference between the contact count pattern ob-
tained for the three slide/roll ratios. The highest occurrence is with
0.25, and the area experiencing the higher rates of asperity
contact is more pronounced for this and the lower sliding case of
0.1. The higher sliding speed of
0.5 seems to lead to a
reduced rate of asperity contact, and with the area over which high
asperity contact rates occur also reduced in comparison with the
other cases. This feature of the results follows from the fact that
the entrainment effect for asperity/asperity collisions within the
Hertzian region is effectively given by 0.5 times the sliding ve-
locity as was demonstrated for line contacts in 12. Thus the
asperity/asperity entrainment effect for
0.5 is five times higher
than for
0.1. This observation would seem to suggest that
higher sliding may be advantageous in preventing asperity/
asperity contact. However, we hasten to add that the analysis pre-
sented here is isothermal so that the detrimental effects of local-
ized heat generation due to thin film or dry contact sliding
between the asperities are not included at this stage.
Figure 5 shows contours of contact count per unit time for the
case of contact between two surfaces having Profile B. The con-
tact behavior of these surfaces is quite different as can be seen by
a comparison with Fig. 4. The mesh points experiencing the high-
est contact counts in Fig. 5 are seen to be located in a very limited
area around the boundary of the Hertzian contact region. The con-
tact counts at these locations are about 25% of the peak values
seen in Fig. 4. In addition there are almost no contact occurrences
in the remainder of the Hertzian area. This is in marked contrast to
Figure 4 where contact is seen to occur over the whole width of
the Hertzian area, which suggests strongly that contact occurs
across the whole Hertzian contact area with some particular as-
perity collisions. There are bands in Fig. 4 where no contact has
occurred but these probably result from a lack of asperity colli-
sions at these locations during the analysis time.
Figure 6 shows contours of contact count per unit time for the
case of contact between two surfaces having Profile A. The peak
contact count level is similar for each of the sliding speeds, but
this case illustrates the strong effect of the asperity/asperity en-
trainment due to the sliding velocity. For the highest sliding speed
the contact count rate is close to zero over most of the Hertzian
region and high values are concentrated at the transverse contact
boundary. For the lower sliding speed contact conditions also oc-
cur on the centerline and in bands over the exit half of the contact.
Figure 7 shows contact count contours for a contact consisting
of Profile A running against Profile C. Comparing this figure
with Figs. 4 and 6 shows contact incidences that are intermediate
between those of the individual surfaces in contact with them-
selves. Contact between these two surfaces takes place approxi-
mately half as frequently as that between two surfaces having
profile C, and with a similar pattern of contact intensity/location.
Figure 8 examines the contact count obtained for Profile C for
0.25 in three different cases. In Fig. 8a the surface is in
contact with a smooth surface. In Fig. 8b it is again in contact
Fig. 4 Contours of contact count rate
Q
Õms for the transient
analysis of two surfaces each having Profile C. The heavy
curve indicates a Hertzian dry contact area.
Table 2 Percentage of transient analysis time for which con-
tact is calculated to occur at one or more mesh points
Surfaces
0.1
0.25
0.5
AandA 382634
B and B 19 20 4.3
CandC 989586
AandC 967755
C and smooth 9.2
& C and smooth 48
64 Õ Vol. 127, JANUARY 2005 Transactions of the ASME
Downloaded 04 Apr 2012 to 131.251.133.27. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

with a smooth surface, but the scale of the roughness of Profile
C has been increased by a factor of &. This is so that the
composite roughness is of the same order as that for contact be-
tween two surfaces having Profile C, which is the case illus-
trated in Fig. 8c. In Fig. 8a contact hardly takes place and the
little that does occur is limited to the transverse contact boundary.
In Fig. 8b there is a six-fold increase in the contact count due to
the higher roughness and this is again concentrated at the trans-
verse contact boundary. For the case of the two rough surfaces, the
maximum contact count at the transverse boundary has increased
by a further factor of three, and bands of contact occurrence
spread across the entire contact width. Although the composite
roughness in the cases shown in Figs. 8b and c are the same,
the relative radius of curvature at the tips of the most aggressive
asperities is smaller in the case of Fig. 8b than in the case of Fig.
8c where they will generally be in contact with surface features
having a larger radius of curvature. However contact is less preva-
lent in Fig. 8b than in Fig. 8c which supports the view that
asperity collision is an important factor in causing contact to oc-
cur. This feature of the results, and the percentage contact times
given in Table 2, point towards the possibility of experimental
verification of the predicted contact effects by measuring frac-
tional contact time using electrical contact resistance. Experiments
to investigate this effect are planned for future work.
Figure 9 shows a photograph of part of the surface of a test disk
taken from the scuffing program reported in 5. The disk is from
an experiment where the contact load was removed at the first
indication of scuffing. The disks are crowned and the Hertzian
contact area is illustrated by the ellipse superimposed on the pho-
tograph. Grinding marks can be clearly seen extending across the
width of the disk but the surface finish has been totally changed in
the scuffed part of the running track. The width of the scuffing
mark is about 25% that of the running track, and its outer edge
corresponds to the transverse limit of the Hertzian contact area.
This pattern has been observed in all the scuffing experiments
utilizing transverse-ground crowned disks. The width of the
scuffing mark is dependent on the rapidity with which load is
removed when scuffing is detected by its characteristic sudden
increase in friction. We suggest that the correspondence between
the position of the scuffing track in Fig. 9 and the location of
predicted high contact counts typically shown in Figs. 4 to 7 is
striking. This indicates strongly that the primary cause of scuffing
in these experiments was the breakdown of the EHL film as a
result of direct contact between the surfaces.
Behavior in the Center of the Contact. Although the results
given above concentrate on the features of mixed lubrication situ-
ations brought about by side flow at the transverse edges of the
elliptical contact, it is interesting to compare the center line (y
0) behavior with that of the corresponding line contacts. Figure
10 shows one such example of the film thickness and pressure
distribution at a particular timestep. The figure includes both the
elliptical contact and equivalent line contact results at the same
timestep. It can be seen that the film thickness behavior is identi-
cal between the two methods and the minor differences in pres-
sure are no greater than inevitably exist in the comparison of these
equivalent smooth surface solutions. This equivalence is found to
be generally the case 3 for low conditions with contacts of this
aspect ratio. This gives a clear demonstration that line contact
transient analyses are a suitable tool for investigating micro-
pitting and scuffing in involute gears, failure occurrences which
are not limited to the edges of the gear face width.
The aspect ratio of the contacts considered up to this point are
4:1, i.e. a/b 0.25. For elliptical contacts that have more adverse
aspect ratios the proximity of the transverse boundaries exerts a
greater influence over the main part of the contact area as might
be expected. Figure 11 compares the contact count contours for
the case of two rough surfaces each having profile C with
Fig. 5 Contours of contact count rate
Q
Õms for the transient
analysis of two surfaces each having Profile B. The heavy
curve indicates a Hertzian dry contact area.
Fig. 6 Contours of contact count rate
Q
Õms for the transient
analysis of two surfaces each having Profile A. The heavy
curve indicates a Hertzian dry contact area.
Journal of Tribology JANUARY 2005, Vol. 127 Õ 65
Downloaded 04 Apr 2012 to 131.251.133.27. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Citations
More filters
Journal ArticleDOI

A Review of Elasto-Hydrodynamic Lubrication Theory

TL;DR: The development and understanding of elastohydrodynamic lubrication (EHL) can be traced back to the beginning of the previous century as mentioned in this paper, but it was not until 1949 that the first real solution of the problem was published.
Journal ArticleDOI

On some aspects of numerical solutions of thin-film and mixed elastohydrodynamic lubrication:

TL;DR: In this article, a set of analyzed sample cases with different computational meshes, different cut-off values for handling contact, on different levels of central film thickness, from several hundred nanometres down to zero.
Journal ArticleDOI

Sixty years of EHL

TL;DR: In this article, the authors describe the rheological properties of lubricants under the very severe conditions present in thin-film elastohydrodynamic lubrication contacts, where the fluid film can break down at asperity conjunctions.
Journal ArticleDOI

Elastohydrodynamic Lubrication: A Gateway to Interfacial Mechanics—Review and Prospect

TL;DR: Elastohydrodynamic Lubrication (EHL) is commonly known as a mode of fluid-film lubrication in which the mechanism of hydrodynamic film formation is enhanced by surface elastic deformation and lubricant viscosity increase due to high pressure as mentioned in this paper.
Journal ArticleDOI

Effect of lubricants on micropitting and wear

TL;DR: In this article, a three-contact disc machine with a central roller in contact with three harder, annular counter-discs (rings) of precisely controlled roughness was used for micropitting.
References
More filters
Journal ArticleDOI

Transient elastohydrodynamic point contact analysis using a new coupled differential deflection method Part 1: Theory and validation

TL;DR: In this paper, a transient analysis technique for point contact elastohydrodynamic (EHL) lubrication problems using coupled elastic and hydrodynamic equations is presented, which is validated under transient conditions by a detailed comparison with published work produced using a different, independent method incorporating a moving roughness feature.
Journal ArticleDOI

Conditions for Scuffing Failure of Ground and Superfinished Steel Disks at High Sliding Speeds Using a Gas Turbine Engine Oil

TL;DR: In this paper, the results of an experimental investigation to compare the scuffing performance of conventionally ground and super-finished hardened steel disks operating at sliding speeds of up to 26 m/s and lubricated with a gas turbine engine oil at a temperature of 100° C.
Journal ArticleDOI

Evaluation of deflection in semi-infinite bodies by a differential method:

TL;DR: In this article, a procedure for evaluating the Laplacian of the deflection of a semi-infinite body subject to pressure loading using suitable quadrature expressions is presented.
Journal ArticleDOI

Elastohydrodynamic film formation at the start-up of the motion

TL;DR: In this paper, an experimental study of elastohydrodynamic (EHD) lubricating film formation during the start-up of motion of a point contact from rest is described.
Journal ArticleDOI

Contact and elastohydrodynamic analysis of worm gears Part 1: Theoretical formulation

TL;DR: In this paper, the authors present the theoretical basis for modeling the contact conditions and elastohydrodynamic lubrication of worm gears, the results of which are presented in Part 2.
Related Papers (5)
Frequently Asked Questions (10)
Q1. What are the contributions mentioned in the paper "Analysis of mixed lubrication effects in simulated gear tooth contacts" ?

The paper presents results obtained using a transient analysis technique for point contact elastohydrodynamic lubrication ( EHL ) problems based on a formulation that couples the elastic and hydrodynamic equations. 

The lubrication mechanism primarily responsible for the protection of gear tooth surfaces from wear and surface distress is elastohydrodynamic lubrication ~EHL!. 

Because of the different speeds of the two surfaces the time taken for both surfaces in the contact to become fully rough is that required for the slowest of the two surfaces to move from the inlet boundary to the exit boundary. 

The width of the scuffing mark is about 25% that of the running track, and its outer edge corresponds to the transverse limit of the Hertzian contact area. 

Two practical problems associated with roughness effects and film thinning in gears are micropitting ~rolling contact fatigue on the scale of surface asperities! 

For elliptical contacts that have more adverse aspect ratios the proximity of the transverse boundaries exerts a greater influence over the main part of the contact area as might be expected. 

The detrimental effect of transverse leakage is not confined to the extreme edge of the contact, but can also occur due to transverse waviness ~i.e., 3D roughness! of the contacting components within the overall contact. 

The identification of contact edges as the location of initial scuffing failure is thus a significant observation indicating that failure of the physical mechanism of EHL is a primary underlying cause of scuffing in gear tooth contacts. 

This feature of the results follows from the fact that the entrainment effect for asperity/asperity collisions within the Hertzian region is effectively given by 0.5 times the sliding velocity as was demonstrated for line contacts in @12#. 

The presence of roughness on both surfaces causes a significant variation in both pressure and film thickness within the contact area, and Fig. 3 illustrates this effect for one particular timestep in the analysis for two rough surfaces having profile ~C! with a slide/ roll ratio of j50.25.