You are on page 1of 189

AN INVESTIGATION OF LOAD-INDEPENDENT POWER LOSSES OF GEAR

SYSTEMS

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy
in the Graduate School of The Ohio State University

By
Satya Seetharaman, B.E., M.S.
Graduate Program in Mechanical Engineering

The Ohio State University


2009

Dissertation Committee:
Ahmet Kahraman, Advisor
Vish Subramaniam
Gary L. Kinzel
Robert A. Siston

ABSTRACT

Physics-based fluid mechanics models are proposed to predict load-independent


(spin) power losses of gear pairs due to oil churning and windage. The oil churning
power loss model is intended to simulate spin losses in dip-lubricated conditions while
the windage power loss model is intended to simulate spin power losses under jetlubrication conditions. The total spin power loss, in either case, is defined as the sum of
(i) power losses associated with the interactions of individual gears with the environment
surrounding the gears, and (ii) power losses due to pumping of the oil or air-oil mixture at
the gear mesh.

Power losses in the first group are modeled through individual

formulations for drag forces induced by the fluid, which is the lubricant in the case of oil
churning power losses and air or air-oil mixture in the case of windage power losses, on a
rotating gear body along its periphery and faces, as well as for eddies formed in the
cavities between adjacent teeth. Gear mesh pocketing/pumping losses are predicted
analytically as the power loss due to squeezing of the fluid as a consequence of volume
contraction of the mesh space between mating gears as they rotate. The pocketing losses
are modeled through means of an incompressible fluid flow approach in the case of oil
churning power losses.

When the gear pairs rotate under windage conditions, a

compressible fluid flow methodology is considered for predicting the pocketing losses.
The power loss models are applied to a family of unity-ratio spur gear pairs to quantify
ii

the individual contributions of each power loss component to the total spin power loss.
The influence of operating conditions, gear geometry parameters and lubricant properties
on spin power loss are also quantified.
The oil churning and windage power loss models are validated through
comparisons to extensive experiments performed on spur gear pairs under dip- and jetlubricated conditions, over wide ranges of gear parameters and operating conditions.
The direct comparisons between model predictions and measurements demonstrate that
the model is indeed capable of predicting the measured spin power loss values as well as
the measured parameter sensitivities reasonably well, reinforcing the possibility of
utilizing the proposed model as a computationally effective design tool for predicting
power losses in geared systems. The spin power loss model is further generalized to
handle the several complex and varying gear configurations and operating conditions
present in an actual manual transmission in order to come up with a transmission spin
power loss model, which when coupled with a transmission mechanical power loss model
and existing bearing power loss prediction methodologies, can predict the total power
loss in a transmission. This transmission power loss model formed by these three power
loss components is validated through comparison to actual power loss measurements
from a six-speed example manual transmission, indicating that the transmission power
loss model can indeed be used for design and product improvement activities.

iii

Dedicated to
Amma, Appa
&
Chithra, Sujatha, Kavita

iv

ACKNOWLEDGEMENTS

I would like to express my heart-felt gratitude to my advisor, Prof. Ahmet


Kahraman, for guiding me throughout this fruitful journey, leading to the completion of
my dissertation.

But for his invaluable foresight, discretion and plenteous

encouragement, it would be hard to fathom the maturity of my dissertation being beyond


just nascent, and in turn, my own growth onto an enriching career in research and letters.
It is often said that one need only put in the effort; one need not worry about the fruits of
ones labors. From him, I have learned to conduct research in an honest and dedicated
manner; I have learned to put in the hard yards without expectations. The fruition of such
an effort has been my development as an individual and engineer, en route to a more
complete being. My genuine appreciation is also due towards the wisdom and guidance
of Prof. Vish Subramaniam, all along the very many stages of my dissertation research.
His patience in listening to my research problem provided me with a platform to voice
my ideas and queries when the roads were hard to navigate. Many thanks are due
towards Prof. Gary Kinzel and Dr. Robert Siston for agreeing to serve on my doctoral
committee and for their searching inquiries and critical views of my dissertation research.
I am grateful to the financial support extended by General Motors Powertrain, Europe,
for sponsoring my dissertation work.

I would like to thank Jonny Harianto and Sam Shon for their tremendous help and
technical expertise, in the areas of programming and experimentation. I have spent many
hours in their office, and my admiration for the nature of their job has only grown tenfold
over these years. Sincere thanks also go to Gary Gardner for putting up with my requests
in correcting uncorrectable holes in test rigs; most appreciation to David Talbot for
stimulating discussions on power losses and assistance with Visual Basic programming.
Thanks are also owed to all my lab mates for putting up with me. To Travis PetryJohnson, Mike Moorhead, Tim Szweda and Hai Xu, many thanks towards your help in
gear pair and transmission testing, and friction power loss calculations. To my friends,
particularly, Vijay Kanagala, Sachit Rao and Satyajit Ambike, while those endless hours
of arguments and articulations over coffee and similar such substances have detracted me
from my immediate goals, I have only one thing to say: you all rock! Live on!
I stand as who I am today because of the love, trust and support of my parents.
My words go only a small way in conveying my feelings towards them; so does this
dissertation, which is dedicated to them. I spend my wakeful hours in eternal debt to
them. Amma and Appa, you never stopped believing in me even when I did: look, I have
come home now. My three sisters, for whom this dissertation is dedicated to as well,
have nursed and guided me towards the twin pedestal of dignity and strength. No matter
what, my love and gratitude towards my sisters can never be broken. These have been
several wonderful years, between my transition from callow youth to patient adult. I
have lost a lot, but in the process, I have gained more than what I bargained for. Here is
to hoping my legs will keep running till they no longer can. At least, till one Boston.
vi

VITA

Feb. 18, 1981 ... Born Tiruchirapalli, India


Sep. 1998 Jun. 2002 .

B.E., Mechanical Engineering, Regional


Engineering College, Tiruchirapalli, India

Aug. 2002 Jun. 2004 .

M.S., Mechanical Engineering, Applied


Mechanics, Iowa State University, Ames, Iowa

Sep. 2004 Mar. 2005.

Graduate Teaching Associate, Dept. of


Mechanical Engineering, The Ohio State
University, Columbus, Ohio
Graduate Research Associate, Gear and Power
Transmission, Research Laboratory, Dept. of
Mechanical Engineering, The Ohio State
University, Columbus, Ohio

Mar. 2005 present

PUBLICATIONS
1. Seetharaman, S. and Kahraman, A., 2009, Oil Churning Power Losses of a Gear
Pair: Model Formulation, ASME Journal of Tribology, 131(2), 022201 (11 pages).
2. Seetharaman, S., Kahraman, A., Moorhead, M. D., Petry-Johnson, T., T., 2009, Oil
Churning Power Losses of a Gear Pair Experiments and Model Validation, ASME
Journal of Tribology, 131(2), 022202 (10 pages).

FIELDS OF STUDY
Major Field: Mechanical Engineering
vii

TABLE OF CONTENTS

Abstract.

Page
ii

Dedication.

iv

Acknowledgments....

Vita...

vii

List of Tables........

xii

List of Figures.......

xiii

Nomenclature

xvi

CHAPTERS:

Introduction.....

1.1 Research Background and Motivation..

1.2 Literature Review......

1.3 Scope and Objectives............

1.4 Dissertation Outline..........

References for Chapter 1............

12

viii

Oil Churning Power Losses of a Gear Pair: Model Formulation.......

16

2.1 Introduction.......

16

2.2 Gear Pair Oil Churning Power Loss Model......

20

2.3 Drag Power Losses....... ...

22

2.3.1

Power Loss due to Drag on the Periphery of a Gear.....

22

2.3.2

Power Loss due to Drag on the Faces of a Gear...

29

2.3.2.1 Laminar Flow Conditions...

30

2.3.2.2 Turbulent Flow Conditions.

33

2.4 Power Loss due to Root Filling....

35

2.5 Oil Pocketing Power Loss.

40

2.5.1

Backlash Flow Area......

44

2.5.2

End Flow Area..

50

2.5.2.1 Calculation of Area Qc 2 .

51

2.5.2.2 Calculation of Area Qt(,1mj) ...

54

2.5.2.3 Calculation of Area Qb( m,1 )j ...

56

2.5.3

Power Loss due to Oil Pocketing..

58

2.6 Example Oil Churning Analysis...

62

2.7 Concluding Remarks.

68

References for Chapter 2....

70

Windage Power Losses of a Gear Pair: Model Formulation......

75

3.1 Introduction.......

75

ix

3.2 Windage Pocketing Power Loss Model........

78

3.3 Windage Drag Power Losses........

89

3.4 Example Windage Power Loss Analysis and Parametric Studies....

91

3.5 Concluding Remarks..

105

References for Chapter 3.... 107


4

Validation of Oil Churning and Windage Power Loss Models...... 108


4.1 Introduction....... 108
4.2 Validation of Gear Pair Oil Churning Model... 110
4.2.1

Test Machine and Oil Churning Test Procedure....... 110

4.2.2

Gear Specimens and Parameters Studied...... 114

4.2.3

Power Loss due to Drag on the Faces of a Gear... 118

4.3 Validation of Windage Power Loss Model....... 126


4.4 Concluding Remarks. 130
References for Chapter 4.... 131
5

Application of Oil Churning Power Loss Model to an Automotive transmission.. 132


5.1 Introduction....... 132
5.2 Validation of Gear Pair Oil Churning Model... 136
5.2.1

Test Machine and Oil Churning Test Procedure....... 138

5.2.2

Gear Specimens and Parameters Studied...... 138

5.2.3

Power Loss due to Drag on the Faces of a Gear... 141

5.3 Validation of Windage Power Loss Model....... 142


5.4 Concluding Remarks. 147
x

5.4.1

Test Machine and Oil Churning Test Procedure....... 147

5.4.2

Gear Specimens and Parameters Studied...... 149

5.5 Concluding Remarks. 153


References for Chapter 5.... 154
6

Conclusions and Recommendations for Future Work.... 156


6.1 Summary....... 156
6.2 Conclusions and Contributions..... 157
6.3 Recommendations for Future Work.......... 160

Bibliography ........ 166

xi

LIST OF TABLES
Table

2.1

Page
The design parameters of the example spur gears .......

xii

63

LIST OF FIGURES
Figure

Page

2.1

Definition of oil churning parameters for a gear pair immersed in oil.....

23

2.2

Definition of geometric parameters associated with root filling power losses.

36

2.3

Illustration of a side view of fluid control volumes of the gear mesh


interface at different rotational positions 0 < m1 < m2 < m3 .

2.4

41

Three-dimensional representation of a control volume showing backlash and


end flow areas ..

43

2.5

Geometry of two gears in mesh at an arbitrary position m .

45

2.6

Definition of the end area at an arbitrary position m .......

46

2.7

Parameters used in calculation of (a) the total tooth cavity area Qc 2 ,(b) the
overlap area Qt(,1mj) and (c) the excluded area Qb( m,1 )j , all at an arbitrary position

m ....
2.8

Effect of (a) temperature T , (b) oil level parameter h , (c) face width b and
(d) gear module m% on total spin power loss PT ...

2.9

65

Components of PT for a gear pair having m% = 2.32 mm and b = 26.7, 19.5


and 14.7 mm at h = 1.0 and 80 C; (a) Pdp , (b) Pdf , (c) Prf , and (d) Pp .....

3.1

52

67

Effect of (a) temperature T , (b) face width b and (c) gear module m% on

Pw .....

xiii

93

3.2

Components of Pw for the gear pairs having m% = 2.32 mm and varying b at

80o C ; (a) windage pocketing power loss Pwp (b) windage drag power
losses Pwd .....
3.3

96

Windage power loss Pwp and its components, Pwp ,e and Pwp ,b , for gear pairs
having m% = 2.32 mm at 80o C ; (a) b = 14.7 mm (b) b = 19.5 mm (c)

b = 26.7 mm. ....


3.4

97

Variation of the end flow area Ae(,1m )j with rotational position for the gear
pairs having (a) m% = 2.32 mm , and (b) m% = 3.95 mm.. 100

3.5

Variation of (a) density, (b) pressure and (c) velocity of control volume
( m)
11
of the gear pair having m% = 2.32 mm with rotational position for (a1,

b1,

c1)

the

end

area

and

(a2,

b2,

c2)

the

backlash

area

( b = 19.5 mm , 80 C )....
3.6

102

Variation of (a) density, (b) pressure and (c) velocity of control volume
( m)
11
of the gear pair having m% = 3.95 mm with rotational position for (a1,

b1,

c1)

the

end

area

and

(a2,

b2,

c2)

the

backlash

area

( b = 19.5 mm , 80o C )........

103

4.1

(a) A view and (b) the layout of the gear efficiency test machine [4.1, 4.2].... 111

4.2

One of the test gear boxes shown in dip-lubrication arrangement [4.1]....... 113

4.3

Two examples of test gears (a) 23-tooth gear with

m% = 3.95 mm and

b = 19.5 mm, and (b) 40-tooth gear with m% = 2.32 mm and b = 19.5 mm
[4.1, 4.2].... 115
4.4

Illustration of oil level parameters [4.2]... 117

4.5

Comparison of predicted to the measured [4.2] PsT for a gear pair (a) 30o C ,
(b) 50o C , (c) 70o C , and (d) 90o C ; m% = 2.32 mm, b = 19.5 mm and

h = 1.0 .. 120

xiv

4.6

Comparison of predicted vs. measured PsT for a gear pair at (a) h = 0.05 ,
(b) h = 0.5 , (c) h = 1.0 , and (d) h = 1.5 ; oil at 80o C (up-in-mesh),

m% = 2.32 mm, and b = 19.5 mm.... 122


4.7

Comparison of predicted to the measured [4.1] PsT for a gear pair having

m% = 2.32 mm and face width values (a) b = 14.7 mm, (b) b = 19.5 mm, and
(c) b = 26.7 mm; oil at 80o C and h = 1.0 ......
4.8

124

Comparison of predicted to the measured [4.2] PsT for a gear pair of face
width b = 19.5 mm and modules (a) m% = 2.32 mm and (b) m% = 3.95 mm; oil
at 80o C and h = 1.0 ..... 125

4.9

Comparison of predicted and measured [4.2] Pw for (a) the 40-tooth gear
pair having m% = 2.32 mm and b = 26.7 mm and (b) 23-tooth gear pair
having m% = 3.95 mm and b = 19.5 mm .......

127

5.1

Flowchart of the transmission power loss computation methodology.....

137

5.2

Example system: A 6-speed manual transmission ....... 143

5.3

Rotating gears on the planes of (a) 6th gear pair, (b) 4th gear pair, (c) 3rd and
5th gear pairs, (d) 2nd gear pair, (e) 1st and reverse gear pairs, and (f) final
drive gear pair .......... 144

5.4

Experimental test setup for measuring transmission power losses of the


example 6-speed transmission.

Safety guards are removed for

demonstration purposes............ 148


5.5

Comparison of measured [5.11] and predicted transmission mechanical


power loss values...... 150

5.6

Comparison of measured [5.11] and predicted transmission spin power loss


values: (a) 2nd gear, (b) 3rd gear, (c) 4th gear, (d) 5th gear, and (e) 6th gear.. 152

xv

NOMENCLATURE

Area

Gear facewidth

Friction drag coefficient

Specific heat

Drag force

Control volume

Immersion depth

Dimensionless immersion parameter

Length parameter along gear faces

Total number of gear rotational increments

Gear rotational position index m [1, M ]

m%

Gear module

Number of teeth of gear

Unit normal vector

Average number of cavities

Power loss

Pressure
xvi

Cavity area, Heat

Universal gas constant

Re

Reynolds number

Gear radius

Torque, temperature or tooth thickness

Time

Free-stream velocity

Internal energy

Volume

Velocity vector

Velocity

Work

Axis parallel to gear face

Axis perpendicular to gear face

Boundary layer thickness

Energy per unit mass

Immersion angle

Ratio of specific heats

Momentum thickness

Viscosity

Angle, tangential direction

Density
xvii

Shear stress

Rotational speed in rpm and rad/s, respectively. =

Flow factor

Stream function

Displacement thickness

Subscripts
b

Base or backlash

Cavity

Drag

End

Face

i = 1, 2

Gear index

Index for control volumes

Adiabatic

Kinematic

Rotational position

Outside

Periphery, pitch, pocketing

rf

Root filling

Start of active profile, shaft

xviii

60

Total

Constant Volume

Windage

Superscripts

Rotational position

Laminar

Turbulent

Radial

Tangential

Wall

xix

CHAPTER 1

INTRODUCTION

1.1 Research Background and Motivation


Gears have found prevalent use in automotive industry, turbines and compressors,
gear pumps and most commonly in vehicle drive trains. Since gears transmit power
through rotational motion at different speeds, torques and direction, power is lost through
dissipation taking place due to friction between geared elements and spin losses
accounted for by the environment surrounding the system.
Power losses experienced by drive trains of passenger vehicles have been one of
the major concerns in automotive powertrain engineering over the past few decades.
Such losses directly impact fuel consumption of the vehicle, helping define how good a
vehicle is in terms of its fuel economy and gas/particulate emission levels. Recent
environmental regulations have made it imperative to look into the emission levels and
fuel economy of geared transmissions, and to seek improvements that might enhance
efficiency and reduce power losses. In tow with this line of thought, the study of power
losses and efficiency of geared transmissions has become an important area of interest, as

dwindling fossil fuel resources has stressed the urgent need to come up with means of
improving efficiency of transmissions.
While fuel consumption alone is a sufficient reason for seeking reduced drivetrain
power losses, there are other supplementary reasons as well. Excessive power losses
within the transmission amount to additional heat generation and higher temperatures,
thus, adversely impacting gear contact fatigue and scuffing failure modes [1.1]. A gear
pair that is more efficient will result in lesser heat generation, and hence translates into
better performance. In addition, the design of the lubrication system as well as quantity
of lubricant within the transmission is also related to the amount of heat generated. A
more efficient transmission will free up the demands on the capacity and the size of the
lubrication system; consequently, the amount and quality of the gearbox lubricant are
also eased with improved efficiency. This in turn reduces the overall weight of the unit,
contributing to further efficiency enhancements and reduction in power loss.
In a geared system, the total power loss is comprised of two groups of losses: (i)
load-dependent (friction induced) mechanical power losses and (ii) load-independent
(viscous) spin losses. Sliding and rolling friction losses at the loaded gear meshes and at
the bearings largely define the load-dependent mechanical power losses.

The total

mechanical loss is then given as the sum of losses from all gear meshes and bearings.
The sliding friction losses are related to the coefficient of friction, normal load and
sliding velocity on the contact surfaces while the rolling friction losses occur due to the
formation of an elastohydrodynamic (EHL) film [1.1]. Meanwhile, load-independent
spin losses are caused by a host of factors including viscous dissipation of bearings and
2

gear oil churning and windage that are present as a result of oil/air drag on the face and
sides of the gears as well as pocketing/squeezing of lubricant in the meshing zone.
Moreover, oil shearing taking place in the synchronizers and in seals are also possible
sources of frictional and viscous dissipation, and can be added to the total transmission
power loss.
Spin loss in geared transmissions are of primary importance, as experimental
results have shown them to account for the bulk of the total power loss at higher speeds
of operation and at light loads. The spin power loss can be broken down into several
components as mentioned above. Drag losses, defined as losses taking place on the
periphery and faces of a gear, are computed separately for each gear of the gear pair and
then summed up to give the total drag loss of a gear pair. Power losses can also take
place due to swirling motion of lubricant in the cavity between adjacent teeth through the
mechanism of root filling and subsequent transportation into the meshing zone.
Pocketing/squeezing losses are defined as losses taking place in the meshing zone when
lubricant is squeezed out of the gear meshes due to the pumping action of gear pairs.
When gear pairs rotate in free air, windage loss takes over as the dominant mode of
power loss, requiring a compressible fluid flow approach to compute such losses.

1.2 Literature Review


While there many published studies on prediction of mechanical power losses of a
single gear pair (see ref. [1.1] for a detailed review of literature on such studies), there are
only a few published studies available on modeling spin losses in gear pairs and in multi3

mesh gearboxes and transmissions. At low speeds of rotation, power losses mostly stem
from friction between gear teeth and oil/air drag acting on the faces and periphery of gear
pairs, resulting in friction and oil churning losses. At higher speeds, spin losses due to
pumping action in the meshing zone and windage also come into prominence. Several
seminal studies [1.2-1.4] focused on the drag torque and heat transfer associated with
circulation and secondary flows induced by rotation of discs submerged in a fluid. The
induced flows were dependent on the geometries of the bladed rotating elements and its
enclosure [1.2]. Subsequent methodologies for estimating churning and windage losses
in gears borrowed heavily from the fundamental principles espoused by the above
referenced studies, i.e. [1.2-1.4], and were mostly empirical adaptations of the same. The
environment inside the gearbox, which can at best be described as a pseudo-single phase
mixture of oil and vapor, and also the difficulties in measurement of gear teeth
temperature, has made the study of spin losses very challenging. As computational fluid
dynamics (CFD) tools are not very practical and not readily available to model such a
complex application, most of the published models or formulae to predict windage and/or
churning losses are based on dimensional analysis of experimental data or approximate
hydrodynamic formulations to characterize the flow of lubricant around rotating gears.
Following the works of Blok [1.5] and Niemann and Lechner [1.6], the first in
situ temperature measurements were made to calculate the real surface temperature at all
points along the line of action of mating gears [1.5], for a heat transfer approach towards
computing power losses. Bones [1.7] carried out churning torque loss measurements for
discs of varying geometries at different immersion depths for various fluids. Using a
combination of von Karmans [1.8] equations and a modified Reynolds number, Bones
4

proposed a methodology to calculate churning losses in gears.

Moreover, he also

proposed the existence of three different regimes, which the churning losses fall under:
laminar, transition and turbulent. Terekhov [1.9] proposed a similar methodology for
single and meshing gears by testing several gear pairs of different geometries under
highly viscous lubricant conditions and low rotational speeds [1.10]. Both Bones and
Terekhov expressed churning loss in terms of a dimensionless churning torque. Luke and
Olver [1.10] carried out a number of experiments to determine churning loss in single and
meshed spur gear pairs. Experimental conditions were similar to that proposed by Bones,
and geometry of the gear pairs and operating conditions were varied to study their
influence on churning torque.
Ariura et al. [1.11] measured losses from jet-lubricated spur gear systems
experimentally. They proposed an analysis of the power required to pump the oil trapped
between mating gears. Akin et al. [1.12, 1.13] analyzed the effect of rotationally induced
windage on the lubricating oil distribution in the space between adjacent gear teeth in
spur gears. The purpose of their study was to provide formulations to study lubricant
fling-off cooling.

They proposed that impingement depth of the oil into the space

between adjacent gear teeth and the point of initial contact was an important aspect in
determining cooling effectiveness. Pechersky and Wittbrodt [1.14] analyzed fluid flow in
the meshing zone between spur gear pairs to assess the magnitude of the fluid velocity,
temperature and pressures that result from meshing gear teeth. By far, this study can be
considered as one of the few computational models available to predict the pumping
action of lubricant trapped in the meshing zone, while it does not extend the formulation
to computation of the power loss due to the pumping action in the meshing zone. Diab et
5

al [1.15] came up with an approximate hydrodynamic model of oil-air squeezing in the


meshing zone of spur and helical gear pairs. They inferred this phenomenon of pumping
the lubricant from the gear mesh to be a substantial source of power loss, especially under
high operating speeds [1.15].
Meanwhile, most of the studies on gear windage have been experimental,
focusing on losses of gears or disks of various sizes rotating in free air and surrounded by
different enclosures. Dawson [1.16] studied the problem of gear windage extensively and
investigated windage losses from isolated spur/helical gears rotating in free air. Several
empirical formulae resulted from this study. Diab et al [1.17] carried out windage-related
experiments with disks and gears of various shapes and sizes and subsequently came up
with predictions based on fluid flow around a rotating gear and also through dimensional
analysis, to characterize windage loss in pinion-gear pairs. Eastwick and Johnson [1.18]
provided an extensive review of studies on gear windage to conclude that the general
solution for reducing power loss due to windage has not yet been well-established. Wild
et al. [1.19] studied the flow between a rotating cylinder and a fixed enclosure by using a
CFD model and through experiments. Similarly, Al-Shibl et al. [1.20] proposed a CFD
model of windage power loss from an enclosed spur gear pair.

1.3 Scope and Objectives


As seen from the literature review, spin power loss models mostly incorporate
either a dimensional analysis based approach or an approximate hydrodynamic/CFDbased methodology. The dimensional analysis approach deduces a lot from experimental
6

observations that are particular to the environment whereas the approximate


computational formulations are intensive in effort, and also do not shed sufficient light
into the physics of spin power losses. When the majority of the factors involved in
computing spin losses are taken into account, the nature of the environment surrounding
the gear pairs in a transmission or gearbox makes it extremely difficult to formulate a
fluid mechanics-based approach. Also, to perform extensive studies based on oil level,
gear geometry parameters, oil inlet temperature, lubricant parameters and operating
conditions requires a spin loss model that will be able to throw insight into the
contributions of operating parameters without compromising on computational
efficiency.
Accordingly, the first objective of this dissertation is to develop a physics-based,
analytical, fluid mechanics model of gear pair spin power losses, which is capable of
quantifying the impact of key system parameters with minimal computational effort.
Specific objectives in this regard are as follows:
Develop physics-based fluid mechanics models for prediction of spin power
losses of single gear pairs that incorporates the effect of key parameters such as
gear geometry, operating conditions and lubricant properties on spin power loss.
These models will account for losses associated with oil churning and air
windage.
Perform detailed parametric studies to identify and rank-order the key
parameters influencing spin power losses of a gear pair.

Since gear pair spin loss models were not previously formulated in an analytical,
physics-based manner to accommodate the effect of the various parameters that influence
them, the modeling of power losses of an entire transmission has also been based mostly
on measurements and empirical formulations. Accordingly, the second objective of this
dissertation is to develop a methodology for predicting the overall power loss of a multimesh gear transmission. Specifically, this dissertation will accomplish the following:
Develop a physics-based transmission mechanical power loss model by
incorporating a generalized power flow formulation, which combines a gear
pair mechanical efficiency formulation proposed by Xu et al. [1.1] and the
bearing mechanical power loss model proposed by Harris [1.21].
Generalize the gear pair spin loss model developed for single gear pairs so as to
handle any gear pair configuration observed in a transmission and incorporate
the same with a bearing viscous loss model [1.21] to predict the total spin loss
of the transmission.
Perform detailed parametric studies to identify and rank-order the key systemlevel parameters influencing total power loss of a transmission.
As is the case with any model, these formulations cannot be used in confidence
unless validated through comparisons to actual experiments. The last objective of this
dissertation is the validation of the above proposed models. Specifically:
Validation of the gear pair oil churning power loss model by comparing
predictions to measurements from another sister study on oil churning power
8

losses, over a wide range of input speed, oil inlet temperature, geometric
parameters, oil level and lubricant conditions [1.22].
Validation of the gear pair windage power loss model by comparing predictions
to measurements from the gear windage experiments of Petry-Johnson et al
[1.23] and to the windage experiments of Dawson [1.16].
Validation of both the spin and mechanical power loss predictions of the
transmission power loss model through comparisons to measurements from a
companion study [1.24] on a six-speed sample manual transmission.

1.4 Dissertation Outline


Chapter 2 details the methodology of oil churning power losses in gear pairs, with
subsections focusing on the physics behind drag loss on the periphery and on the faces of
the gears in the laminar and turbulent regime, root filling mode of power loss due to the
formation of eddies in the cavity between adjacent gear teeth, and, finally, power loss due
to pocketing/squeezing taking place in the meshing zone as a consequence of successive
compression/expansion of tooth cavity volume.

The oil churning power loss

formulations assume the fluid surrounding the gears as well as the fluid inside the
meshing zone to be incompressible in nature. Detailed formulations are presented for
each component of oil churning power loss, with the chapter winding up with an example
oil churning analysis that includes extensive parametric studies on the influence of

lubricant parameters, gear geometry parameters and operating conditions on oil churning
power losses.
Chapter 3 is devoted to the formulation of a gear pair windage power loss model.
The nature of the compressible air-oil mixture in the meshing zone under jet-lubrication
conditions affords for a departure from the incompressible pocketing power loss
formulation developed as a part of the oil churning power loss model. A modeling
strategy based on the conservation laws is proposed to calculate the power loss due to the
compressible mixture being squeezed out through the meshing zone; drag loss
formulations from the oil churning power loss model are adapted to include the effect of
windage conditions. Chapter 3 also presents an example windage power loss analysis,
with adequate light thrown on the influence of system parameters on the total windage
power loss, which will be given by the sum of the windage drag and windage pocketing
power losses.
In Chapter 4, the gear pair oil churning and windage power loss models delineated
in Chapters 2 and 3 are validated through comparisons to measurements conducted on a
family of unity-ratio spur gear pairs, under dip- and jet-lubricated conditions, so as to
simulate oil churning and windage modes of operation. This extensive validation effort
includes measurements conducted at different speed and temperature conditions, over
different gear design variations as well as with the gear pairs being operated under
different lubricant conditions.
Chapter 5 deals with the proposition and validation of the transmission power loss
model, which incorporates the generalized transmission spin and mechanical power loss
10

models, along with an existing bearing power loss formulation, to compute the total
power loss in an automotive transmission. Validation efforts are carried out through
comparisons to measurement from a sample six-speed manual transmission operated over
a wide range of parameters and under several gear configurations.
Literature review pertaining to particular parts of the formulation proposed in this
dissertation is made available at the beginning of each chapter towards an easy
understanding of the body of work lying behind the relevant models. The repetition of
such critical reviews on initial and existing methodologies between the introductory
chapter and the rest of this dissertation is intentional.
Finally, Chapter 6 provides an extended summary of the entire work while
outlining the major conclusions and contributions of this dissertation. Also, a detailed list
of recommendations for future work is presented at the end of Chapter 6.

11

References for Chapter 1


[1.1]

Xu, H., Kahraman, A., Anderson, N.E., Maddock, D.G., 2007, Prediction of
Mechanical Efficiency of Parallel-Axis Gear pairs, ASME Journal of Mechanical
Design, 129, 58-68.

[1.2]

Daily, J. W., and Nece, R. E., 1960, Chamber Dimensional Effects on Induced
Flow and Frictional Resistance of Enclosed Rotating Disks, ASME Journal of
Basic Engineering, 82, 217232.

[1.3]

Mann, R.W., and Marston, C.H., 1961, Friction Drag on Bladed disks in
Housings as a Function of Reynolds Number, Axial and Radial Clearance and
Blade Aspect Ratio and Solidity, ASME Journal of Basic Engineering, 83 (4),
719-723.

[1.4]

Soo, S. L., and Princeton, N. J., Laminar Flow Over an Enclosed Rotating Disc,
Transactions of the American Society of Mechanical Engineers, 80, 287-296.

[1.5]

Blok, H., 1957, Measurement of Temperature Flashes on Gear Teeth under


Extreme Pressure Conditions, Proceeding of The Institution of Mechanical
Engineers, 2, 222-235.

[1.6]

Niemann G., and Lechner G., 1965, The Measurement of Surface Temperature
on Gear Teeth, ASME Journal of Basic Engineering, 11, 641-651.

12

[1.7]

Bones, R.J., 1989, Churning Losses of Discs and Gears Running Partially
Submerged in Oil, Proceedings of the ASME 5th International Power
Transmission and Gearing Conference, Chicago, 355-359.

[1.8]

Von Karman, T., 1921, On Laminar and Turbulent Friction, Z. Angew. Math.
Mech., 1, 235-236.

[1.9]

Terekhov, A.S., 1975, Basic Problems of Heat Calculation of Gear Reducers,


JSME International Conference on Motion and Powertransmissions, Nov. 23-26,
1991, 490-495.

[1.10] Luke, P., and Olver, A., 1999, A Study of Churning Losses in Dip-Lubricated
Spur Gears, Proc. Inst. Mech. Eng.: J. Aerospace Eng., Part G, 213, 337346.
[1.11] Ariura, Y., Ueno, T., and Sunamoto, S., 1973, The lubricant churning loss in
spur gear systems, Bulletin of the JSME, 16, 881-890.
[1.12] Akin, L. S., and Mross, J. J., 1975, Theory for the Effect of Windage on the
Lubricant Flow in the Tooth Spaces of Spur Gears, ASME Journal of
Engineering for Industry, 97, 12661273.
[1.13] Akin, L. S., Townsend, J. P., and Mross, J. J., 1975, Study of lubricant jet flow
phenomenon in spur gears, Journal of Lubrication Technology, 97, 288-295.
[1.14] Pechersky, M. J., and Wittbrodt, M. J., 1989, An analysis of fluid flow between
meshing spur gear teeth, Proceedings of the ASME 5th International Power
Transmission and Gearing Conference, Chicago, 335342.
[1.15] Diab, Y., Ville, F., Houjoh, H., Sainsot, P., and Velex, P., 2005, Experimental
and Numerical Investigations on the Air-Pumping Phenomenon in High-Speed

13

Spur and Helical Gears, Proceedings of the Institution of Mechanical Engineers,


Part C: J. Mechanical Engineering Science, 219, 785-800.
[1.16] Dawson, P. H., 1984, Windage Loss in Larger High-Speed Gears, Proceedings
of the Institution of Mechanical Engineers, Part A: Power and Process
Engineering, 198(1), 5159.
[1.17] Diab, Y., Ville, F., and Velex, P., 2006, Investigations on Power Losses in High
Speed Gears, Journal of Engineering Tribology, 220, 191298.
[1.18] Eastwick, C. N., and Johnson, G., 2008, Gear Windage: A Review, ASME
Journal of Mechanical Design, 130, 034001, 6 pages.
[1.19] Wild, P. M., Dijlali, N., and Vickers, G. W., 1996, Experimental and
Computational Assessment of Windage Losses in Rotating Machinery, ASME
Trans. J. Fluids Eng., 118, 116122.
[1.20] Al-Shibl, K., Simmons, K., and Eastwick, C. N., 2007 Modeling Gear Windage
Power Loss From an Enclosed Spur Gears, Proceedings of the Institution of
Mechanical Engineers, Part A, 221(3), 331341.
[1.21] Harris, A. T., 2001, Rolling Bearing Analysis, Fourth Edition, Wiley & Sons, Inc.,
New York.
[1.22] Moorhead, M., 2007, Experimental Investigation of Spur Gear Efficiency and
the Development of a Helical Gear Efficiency Test Machine, M.S. Thesis, The
Ohio State University, Columbus, Ohio.
[1.23] Petry-Johnson, T. T., Kahraman, A., Anderson, N.E., and Chase, D. R., 2008, An
Experimental Investigation of Spur Gear Efficiency, ASME Journal of
Mechanical Design, 130 (6), 062601, 10 pages.
14

[1.24] Szweda, T, A., An Experimental Study of Power Loss of an Automotive Manual


Transmission, MS thesis, The Ohio State University, Columbus, Ohio, 2008.

15

CHAPTER 2

OIL CHURNING POWER LOSSES OF A GEAR PAIR: MODEL


FORMULATION

2.1 Introduction
Power losses in gear pairs and transmissions can be broadly classified into two
groups as detailed in the first chapter: load-dependent (friction induced or mechanical)
power losses and load-independent (spin) power losses due to viscous dissipation. In
automotive applications, geared components of systems such as manual transmissions,
transfer cases and front or rear axles might rotate at reasonably high speeds (say, gear
pitch-line velocities in excess of 20 to 30 m/s) to cause significant amounts of spin power
losses. While load-dependent and spin power losses can be comparable in magnitude
under high-load and low operating speed conditions, the spin losses typically dominate
the overall power losses at these higher operating speeds.
Spin power losses of a gearbox are either due to churning of the lubricant if the
rotating components of the gearbox are immersed in an oil bath (dip-lubricated) or due to
windage if the lubrication method is jet-type and the surrounding medium is air or a fine
mist of air and oil. Focusing on the most fundamental component of the gearbox, i.e. a
gear pair in mesh, this chapter aims at developing a novel physics-based fluid
16

mechanics model of oil churning losses due to interactions of the gears, both as a pair and
as individual entities, with the surrounding lubricant medium. A companion model to
handle windage losses in a jet-lubrication scenario will be proposed in Chapter 3.
While there has been a large body of work dealing with load-dependent power
losses ([2.1, 2.2]), there are only a few published studies on modeling spin losses.
Several studies (e.g. [2.3-2.4]) proposed formulations for the drag torque associated with
circulation and secondary flows induced by rotation of a disc submerged in a fluid.
Subsequent studies on churning and windage losses in gears relied heavily on these
fundamental studies, and they can mostly be characterized as empirical adaptations of the
same. The convolution of the environment inside the gearbox, which can at best be
described as a pseudo-single phase mixture of oil and vapor, as well as the complexity of
gears rotating in mesh have made the study of spin losses a very challenging one. As
computational fluid dynamics (CFD) tools are not very practical and not readily available
to model such a complex application, most of the published models or formulae to predict
windage and/or churning losses are based on dimensional analysis of experiments or
approximate hydrodynamic formulations to characterize the flow of lubricant around
rotating gears.
Taking a critical view of existing methodologies, Bones [2.5] carried out churning
torque measurements for discs of varying geometries at different immersion depths using
three different fluids. Using a combination of von Karmans [2.6] equations and a
Reynolds number based on the disc chord length as the characteristic length, he proposed
a methodology to calculate churning losses in gears in three different regimes: laminar,
17

transition and turbulent. Terekhov [2.7] proposed a similar methodology for single and
meshing gears based on gear experiments under highly viscous lubricant conditions and
low rotational speeds. Both Bones and Terekhov expressed churning loss in terms of a
dimensionless churning torque.

Luke and Olver [2.8] performed a number of

experiments to determine churning loss in single and meshed spur gear pairs. They
compared their experimental observations on spin power losses with the empirical
formulations of Bones [2.5] and Terekhov [2.7] and found that contrary to what Bones
had predicted, the spin power losses were not strongly affected by the viscosity of the
lubricant.

Furthermore, their observations called into question the attempt used to

characterize spin power loss based on a Reynolds number dependent on lubricant


viscosity. Ariura et al. [2.9] measured losses from jet-lubricated spur gear systems
experimentally. They proposed an analysis of the power required to pump the oil trapped
between mating gears. Akin et al. [2.10, 2.11] analyzed the effect of rotationally induced
windage on the lubricating oil distribution in the space between adjacent gear teeth in
spur gears. The purpose of their study was to provide formulations to study lubricant
fling-off cooling.

They proposed that impingement depth of the oil into the space

between adjacent gear teeth and the point of initial contact was an important aspect in
determining cooling effectiveness.
Pechersky and Wittbrodt [2.12] analyzed fluid flow in the meshing zone between
spur gear pairs to assess the magnitude of the fluid velocity, temperature and pressures
that result from meshing gear teeth. This work can be considered as one of the few
computational models available to study the pumping action of lubricant trapped in the
meshing zone, while it does not extend the formulation to computation of the power loss.
18

Diab et al [2.13] came up with an approximate hydrodynamic model of oil-air squeezing


in the meshing zone of spur and helical gear pairs. They inferred this phenomenon of
pumping the lubricant from the gear mesh to be a substantial source of power loss,
especially under high operating speeds [2.13]. A more recent study by Changenet and
Velex [2.14] investigated the influence of meshing gear on oil churning power losses by
performing a number of gear oil churning experiments to come up with empirical
formulae for power losses. Parameters included were gear module, diameter and face
width, speed and lubricant viscosity.

Their empirical formulae suggested that the

influence of viscosity on oil churning loses was insignificant at high speeds of rotation
for single gears, corroborating similar findings from the experimental observations of
Luke and Olver [2.8]. Another relevant work by Hhn et al [2.15] also stresses this
apparent lack of dependence of oil type on load independent losses. In their experiments,
Hhn et al [2.15] measured gear and bearing power losses and forged a balance between
generated heat in the gearbox due to gears and bearings and the dissipated heat in the
form of free and forced convection and through radiation as well, from housing and
rotating parts, to calculate mean lubricant temperature.
Examination of the above body of literature reveals that gear pair spin power loss
models were either based on dimensional analysis of controlled experimental data or a
CFD-based computational methodology, with unconvincing light thrown on the influence
of key system parameters on the viscous power losses. Both approaches have major
limitations in terms of their practicality and computational demand, and hence, come
short of providing a better understanding of gear spin power losses, including the
significance of key system parameters. Accordingly, this chapter aims at developing a
19

physics-based, analytical, fluid mechanics model of gear pair spin power losses, which is
capable of quantifying the impact of key system parameters with minimal computational
effort. Load-independent windage power losses will be handled separately in Chapter 3
by taking into account the compressible nature of air or air-oil mixture surrounding the
gears and in the meshing zone, with validation of both the models presented in Chapter 4
through comparisons to tightly controlled experiments and empirical studies.

2.2 Gear Pair Oil Churning Power Loss Model


The churning power losses of a gear pair can be grouped into two categories. The
first category is comprised of drag losses associated with the interactions of each
individual gear with the surrounding medium. For a gear i that is partially or fully
immersed in oil, the drag power loss will be modeled as the sum of three individual
components:

Pdi = Pdpi + Pdfi + Prfi ,

i = 1, 2,

(2.1)

where Pdpi is the power loss due to oil/air drag on the periphery (circumference) of a
gear, Pdfi is the power loss due to oil/air drag on the faces (sides) of a gear, and Prfi is
the power loss that occurs during the filling of the cavity between adjacent teeth with oil.
With these components determined for each gear, the total drag power loss of the gear
pair is then found as Pd = Pd 1 + Pd 2 .

20

The second category of churning power losses consists of losses due to


interactions of the gear pair with surrounding medium (oil) at the gear mesh interface,
with squeezing/pocketing losses being the dominant mode of power loss. With the gear
mesh pocketing power losses Pp and drag power losses predicted, the total churning
power losses of a gear pair is then found as

PT = Pd + Pp .

(2.2)

The following sections describe the details of the oil churning power loss model.
Specifically, Section 2.3 provides the drag power loss formulations along the periphery
and faces of each gear and Section 2.4 details the power loss due to pocketing,
establishing the framework for computing the backlash and end flow areas, as well as
computing the power loss due to pocketing. Many other sources of churning power
losses can also be identified, including those associated with lift-off of the leftover oil
within the mesh at the exit of the mesh and power loss due to the transport or acceleration
of the lubricant swirling inside each tooth cavity as it is carried into the meshing zone.
These effects are not included in the proposed model as preliminary formulations of these
effects suggest that they are secondary. Similarly, the effect of enclosures in the form of
flanges or shrouds in the near vicinity of rotating gears as reported experimentally by
Changenet and Velex [2.16] will not be included as such effects are beyond the scope of
this dissertation.

21

2.3 Drag Power Losses


2.3.1 Power Loss Due to Drag on the Periphery of a Gear
A rotating gear pair that is fully or partially immersed in oil, as shown in Figure
2.1, is subjected to drag forces that are induced along the direction of flow on the
periphery (circumference) and faces (sides) of the gears, thus, contributing to churning
power losses. A gear pair rotating in free air experiences similar drag forces in the form
of air windage, as detailed in Chapter 3. In formulating the drag forces and drag power
losses due to oil churning, each gear is modeled as an equivalent circular cylinder of
radius roi (the outside radius of the gear). This employs the assumption that at medium
to high speeds of rotation, the behavior of a gear immersed in oil follows that of a
cylindrical disk, as the oil swirling around the gear will not feel the effects of the tooth
cavities.

The implication of the above assumption can be extended to gears with

increased tooth height as well, because the Reynolds number defined below in the
following section takes into account the characteristic length scale as the outside diameter
of the gears and an increase in tooth height will be reflected in an increase in the outside
diameter and hence can be accommodated in the drag loss formulation. It has to be noted
though that at low speeds of rotation, substantial changes in tooth thickness will change
the nature of the boundary layer from that seen for flow around a circular disk.
Steady-flow conditions are assumed in this formulation such that v t = 0 and

p t = 0 .

The oil pressure is assumed to vary only in the radial direction, i.e.

pi = pi (r ) and pi = 0 .
22

Gear 1

1
Gear 2

l1

O1

Oil Level

l2

ro1

ro 2

h1

O2

Figure 2.1 Definition of oil churning parameters for a gear pair immersed in oil.

23

h2

(r )
Further, the radial component of the oil velocity vdpi
is considered to be zero. In

addition, flow conditions are assumed to be incompressible, with density remaining


unaffected by changes in pressure, while the viscosity of the lubricant has a strong
exponential dependence on the operating temperature. It is further assumed that under
incompressible flow conditions, changes in pressure in the meshing zone have a
negligible effect on the operating viscosity and density of the lubricant. Also, changes in
pressure close to the contact zone are not relevant to the load-independent power losses,
and as a result, overall changes in pressure are not reflected in the viscosity
characteristics of the lubricant, either in the meshing zone or the contact region.
Meanwhile, the bulk temperature of oil was assumed to be known in this model.
From the standpoint of the analysis, the drag along the periphery reflects the case
of flow in the annulus between two rotating disks, with the outside disk remaining
stationary and placed at infinity to simulate the casing and the inner disk rotating at the
operating speed of the gears in consideration. Regarding the presence of any flanges
enveloping the gears, the drag formulations are valid for as long as the distance between
the flanges or shrouds and the gear surfaces is greater than the total boundary layer
thickness of both the surfaces, i.e. from the gears and the flanges, and as such, the current
formulation cannot, in its entirety, be extended to include the effect of flanges. A more
qualitative analysis, taking into account the interactions between boundary layers arising
from the surfaces of the gear as well as that of the flanges, once the flanges are close
enough to the gear surfaces, must be taken into account to model the presence of flanges.
This boundary layer could be a combined boundary layer or a separate boundary layer,
24

under laminar or turbulent conditions, based on the rotational speed. Future formulations
and extensions to the drag power losses will need to focus on the same. Also, surface
roughness effects are not taken into consideration while formulating the drag power
losses. Surface roughness might play a tertiary role in case of gears with very rough
surfaces, potentially changing the behavior of drag forces, especially in the turbulent
regime. The drag formulations consider only friction drag and the effects of form drag
has not been characterized in this dissertation. The inclusion of form drag will need more
information on flow separation and behavior of the lubricant beyond the wake region and
as has been noted earlier, it is assumed that there is a negligible wake region on
separation, which gives enough reason to neglect form drag in the current formulation.
Also to be noted is that the effects of air ingestion at medium speeds and low oil levels
will change the nature of the analysis from incompressible to compressible flow and
hence, the dependence of density on temperature and pressure would change, bordering
the analysis on two-phase flow.

Imposing single-phase conditions on the drag

formulations can be considered as a gross simplification that avoids the computational


difficulties involved in a two-phase mixture flow.
( )
of the oil along the
Based on the above detailed assumptions, the velocity vdpi

tangential direction on the periphery of the gear and the pressure distribution around the
periphery are defined by solving the continuity equation [2.17]

(r )
vdpi

(r )
vdpi

()

1 vdpi
+
+ [
]= 0,
r
r
r

(2.3a)

25

and the Navier-Stokes equations of motion [2.17] along r and directions

(r )
vdpi

( )
vdpi

r)
+ v (dpi

( )
+ vdpi

(r )
vdpi

( )
vdpi

()
(r )
vdpi
vdpi

( r ) ()
vdpi
vdpi

1 pi

+ 2 [r
r r

() 2
[vdpi
]

=
r
()
2 (r )
vdpi
1 pi
vdpi
(r )
[
],

+
vdpi 2
r r 2 2

( )
( )
vdpi
vdpi

( )
2vdpi
2

(2.3b)

=
()
2vdpi

+r

( )
vdpi

+2

(r )
vdpi

(2.3c)
( )
vdp
].

where and are the density and the dynamic viscosity of the lubricant at a given
operating temperature. Implementing the assumptions stated earlier, one obtains

()
vdpi

= 0,

(2.3d)

( ) 2 pi
[v ] =
,
r dpi
r

(2.3e)

()
d 2vdpi

dr 2

()
( )
1 dvdpi vdpi
+
2 = 0.
r dr
r

(2.3f)

The boundary conditions for the oil circulating along the periphery of the gear (and also
at the end of the enclosure, assuming the gear to be immersed in an enclosure of infinite
( )
( )
= i roi at r = roi and vdpi
= 0 as r , where i is the
depth) are given as vdpi

26

rotational speed of the gear and r [roi , ) . Solving Eq. (2.3f) and applying these
boundary conditions, the fluid velocity along the periphery of the gear is found as

( )
vdpi

i roi2
=
.
r

(2.4)

Given the dynamic viscosity of the oil, the radial, axial and tangential
components of the shear stress built up on the periphery due to the rotational motion are
evaluated next as

(r )
vdpi

rr )
(dpi

= 2

=0,

(2.5a)

)
(dpi

()
(r )
1 vdpi vdpi
= 2 [
+
] = 0,
r
r

(2.5b)

r )
(dpi

2i roi2
v dpi 1 v dpi
= [r (
)+
]=
.
r r
r
r2

()

(r )

(2.5c)

The above set of equations indicates that only the tangential shear stress acting on the
periphery of the gear is nonzero. From Eq. (2.5c), the tangential shear stress at the
w)
= 2i .
outside radius of gear i, r = roi (tangential wall shear stress), is found as (dpi
w)
Defining the friction drag coefficient as Cdpi = 2(dpi
U i2 , where U i = i roi is the free-

27

stream velocity of the lubricant near the periphery of the gear i, the drag force acting on
the periphery of the gear is found as

1
w)
Fdpi = U i2 Adpi Cdpi = Adpi (dpi
.
2

(2.6)

Here, the wetted surface area of the periphery is given as Adpi = 2i roi bi , where bi is face
width of gear i and i = cos 1[1 hi ] . A dimensionless immersion parameter is defined
from Figure 2.1 as hi = hi roi , where hi is the immersion depth of gear i. Accordingly,
hi 0 represents the case of no oil-gear interactions and hi 2 corresponds to the case
of a gear that is submerged in oil. For hi 2 , hi = 2 is used in the above formulation to
obtain i = such that the entire periphery of the gear is subject to drag, i.e.

Adpi = 2roi bi . For hi 0 , it is understood that the gear is submerged in air. Then, the
parameters hi = 2 and i = are used together with the properties of air to calculate air
windage loss at the periphery of the gear, as proposed in Chapter 3.
It has to be noted that the oil levels considered are static oil levels, though in
reality, oil levels will not remain static with gear rotations, resulting in dynamically
changing oil levels. Inclusion of these dynamic oil levels that would require a more
advanced computational model is beyond the scope of the present formulation whose
main aim is to provide a simple, physics-based explanation of the major phenomena
pertinent to oil churning losses. Also, to a certain extent, the assumption of a static oil

28

level is still relevant based on the fact that a reduction of oil level at one chamber or
location will average off the increase in dynamic oil levels at some other location within
the transmission/gearbox. As a result, the averaging effect of dynamic oil level plays into
the hands of assuming static oil levels to circumvent modeling difficulties. With Adpi
w)
and (dpi
known, the drag force is then evaluated as

Fdpi = 4 bi roi i i .

(2.7a)

Finally, the power loss due to drag on the periphery of a single gear is given as the
product of Fdpi and the tangential velocity at the periphery as

Pdpi = 4 bi roi2 i2 i .

(2.7b)

2.3.2 Power Loss Due to Drag on the Faces of a Gear


In modeling the power loss due to drag on the faces of a gear, laminar and
turbulent flow regimes will be handled separately. At low to medium speeds, depending
on the size of the individual gears, the flow can be assumed to be laminar, with transition
to turbulence taking place at higher rotational speeds corresponding to a Reynolds
number defined as Re = 2i roi2 within the range (10)5 to (10)6 [2.17]. Because of a
large velocity gradient between outer and inner flow due to the no-slip boundary
conditions on the surface of the gear, viscous effects are no longer negligible and a
29

boundary layer is formed along the faces of the gear, in which viscous effects become
predominant. Assuming that the boundary layer remains close to the surface of the gear,
with a negligible wake region formed on separation, the flow across the faces of a gear
can be modeled as flow across a flat plate, under laminar or turbulent flow conditions. In
cases when the gear is fully submerged in oil, an alternate method of flow near a rotating
disk [2.18] can be used, which provides the same dependence of power loss on the faces
with the operating parameters as detailed below in the present formulation. However, the
approach of modeling the flow across the faces as flat plate flow is more advantageous
since it can handle partial immersion cases and turbulence more conveniently.

2.3.2.1 Laminar Flow Conditions: In the laminar regime, the flow velocity
( L)
profile vdfi
is considered to be linear, i.e.

( L)
vdfi
= U i y i( L ) ,

(2.8a)

where i( L ) is the laminar boundary layer thickness, U i is the free-stream velocity of the
lubricant flowing across the faces of the gear i as defined previously, and y is the axial
direction perpendicular to the faces (coming out of the paper) along which the velocity
profile varies. The displacement thickness and momentum thickness of the boundary
layer are defined, respectively, as

30

dfi

i =

[1

0
dfi

i =

vdfi
Ui

vdfi
Ui

[1

] dy = 12 i ,

vdfi
Ui

(2.8b)

] dy = 16 i .

(2.8c)

Here, i gives the distance that the outer streamlines are shifted or displaced outward of
the surface of the gear as a result of retarded flow in the boundary layer, whereas i
represents the reduced momentum flux in the boundary layer as a result of shear force on
the surface of the gear [2.17]. Due to the mostly viscous flow across the gear surface,
with k being the kinematic viscosity of the oil, the skin friction coefficient, Ci( L ) must
be taken into account and is defined as

Ci( L ) =

( L)
2 dvdfi
U i2 dy

=
y =0

2 k
( L)
i U i

(2.8d)

The boundary layer momentum integral equation of von Karman [2.17] is applied to
solve for the boundary layer thickness. Since flow across the faces of the gear is taken to
be the case as flow across a flat plate for which the free-stream velocity U i is constant,
the rate of increase of momentum thickness is directly proportional to the wall shear
stress, i.e.

31

i 1
= Ci .
x 2

(2.9)

By setting Ci = Ci( L ) and i = i( L ) to represent the laminar regime and incorporating


Eqs. (2.8b) and (2.8c), Eq. (2.9) is solved to obtain the boundary layer thickness as

i( L ) = 3.46

k x
,
Ui

(2.10)

where x is the length parameter, as defined in Figure 2.1. At x = l i = 2roi sin i , the skin
friction coefficient is obtained by substituting Eq. (2.10) into Eq. (2.8d) as

Cl( L ) = 0.578
i

k
.
li Ui

(2.11)

The drag force on the face of the gear in the laminar regime is then given in terms of the
skin friction coefficient at x = l i as

1
( L)
Fdfi
= U i2 Adfi Cl( L ) .
i
2

(2.12a)

The wetted area of the face is

32

Adfi = roi2 [ 2 sin 1 (1 hi ) (1 hi ) hi (2 hi )] .

(2.12b)

For the case of a fully submerged gear (or for air windage), hi = 2 and the wetted surface
area of the gear reduces to Adfi = roi2 while the length parameter is x = l i = 2roi . Eq.
(2.12a) takes into account only one face of the gear while calculating the drag force.
Hence, in order to include both faces of the gear, Eq. (2.12a) is multiplied by a factor of 2
and further simplified to obtain

( L)
Fdfi

1.5
0.410.5
k U i Adfi

roi sin i

(2.12c)

With U i = i roi , the power loss due to drag on the faces of the gear in the laminar regime
is given as the product of the drag force on the faces of the gear and the tangential
velocity of the gear as

( L)
Pdfi

2.5 2
0.41 0.5
k i roi Adfi

sin i

(2.13)

2.3.2.2 Turbulent Flow Conditions: For turbulent flow over a flat plate, the
velocity profile obeys the Prandtl one-seventh power law [2.17]

33

(T )
vdfi
= Ui [

y
i(T )

]1/ 7 ,

(2.14a)

where the superscript T indicates turbulent flow, with i(T ) being the boundary layer
thickness in the turbulent regime. The displacement and momentum thicknesses are
found from Eqs. (2.8b) and (2.8c), with the velocity profile defined in Eq. (2.14a) as

i(T ) = 18 i(T ) and

7 (T ) . Defining the skin friction coefficient for flow over a


i(T ) = 72
i

rough flat plate that is turbulent from the leading edge as [2.17]

Ci(T ) =0.02[

k
U i i(T )

]0.167 ,

(2.14b)

and substituting the expressions for Ci(T ) , i = i(T ) and i = i(T ) into the momentum
integral equation, the boundary layer thickness is derived as

T)
(dfi
= 0.142 x6 / 7 (

k 1/ 7
) .
Ui

(2.15a)

As in the laminar flow formulation, the value of Ci(T ) at x = l i , as shown in Figure 2.1,
is derived from Eq. 2.14(b) as

Cl(T ) = 0.0276 [
i

k 1/ 7
] ,
l iU i

(2.15b)

34

and the drag force on the faces of the gear under turbulent flow conditions is given as

(T )
Fdfi
= 12 U i2 Adfi Cl(T ) .

(2.16a)

Substituting Eq. (2.15b) and accounting for the both faces of the gear, the drag force on
the faces of the gear i in the turbulent flow regime is written as

(T )
Fdfi
=

1.86 1.72
0.0250.14
roi Adfi
k i

(sin i )0.14

(2.16b)

Finally, the power loss due to drag on the faces of the gear in the turbulent flow regime is
found as

(T )
Pdfi
=

2.86 2.72
0.025 0.14
roi Adfi
k i

(sin i )0.14

(2.17)

2.4 Power Loss due to Root Filling


When a gear pair is partially immersed in oil, the space (cavity) between adjacent
teeth is filled with oil as it enters the oil bath, with the number of filled cavities
depending on the immersion depth of the gears. The rate of filling of the cavities
depends on speed of rotation and the operating temperature. The cavities will be filled
completely with lubricant when the operating temperature is high (i.e. the oil viscosity is
35

low) while partial filling could take place at lower operating temperatures. When the oil
swirls past the edges of a tooth cavity, it can be modeled as flow across an annular cavity.
This swirling action of the oil, caused by the tooth surfaces acting as the sidewalls,
creates eddies within the cavity. The flow direction of eddies differ from that of the
general flow, causing energy dissipation and power loss.
Figure 2.2 shows a schematic of flow across an annular cavity. The sidewalls
representing the inner sides of the gear teeth are considered stationary for this model. It is
assumed that the gear is stationary and the lubricant is swirling across the adjacent
cavities. Further, the flow is assumed to be a creeping flow of an incompressible viscous
liquid (oil). When the flow is relatively slow, the terms involving squares of velocity in
the Navier-Stokes equation can be neglected, allowing for an analytical solution.
Denoting the stream function by (r , ) , the following assumptions are made:
(i) Flow is assumed to be radial with all ( ) = 0 and i (r , ) = i (r ) .

(ii) Steady Stokes flow is assumed, i.e. 4i (r ) = 0 .


(iii) Only radial solutions are considered for the bi-harmonic equation,
(iv) Any effect of gravity is neglected.
In addition, as the exact geometry of the cavity between two teeth shown in
Figure 2.2(b) is rather complex, an approximate annular shape is considered in its place.
The angle of this approximate annular shape, 2ci , is defined by the angle of the cavity at

36

Oi
(a)

rbi

bi
Oil

roi

rri

i
n =1

n = ni
n=2

n = ni 1

(b)

dr

d ci

2ci

roi

rpi
rri
Figure 2.2 Definition of geometric parameters associated with root filling power losses.
37

the pitch radius, rpi , and the outer and inner boundaries are formed by the outside
diameter, roi , and the root radius, rri , respectively, as shown in Figure 2.2(b).
Since the flow across the cavity is assumed to be a creeping flow caused by
tangential velocities on the top and bottom curved boundaries, the bi-harmonic equation
for radial flow can be written as

4i =

d 4 i
dr 4

2 d 3 i 1 d 2 i 1 d i

+
=0,
r dr 3 r 2 dr 2 r 3 dr

(2.18a)

whose solution has the form

i = D1i + D2i r 2 + D3i log r + D4i r 2 log r .

(2.18b)

Given the boundary conditions

0 at r = roi
i =
0 at r = rri
d (i )
i roi
=
dr
i rri

ci ,

(2.19a)

ci ,

at r = roi
at r = rri

ci ,
ci ,

one obtains from Eq. (2.18b)

38

(2.19b)

D1i + roi2 D2i + log roi D3i + roi2 log roi D4i = 0 ,
2roi D2i +

1
D3i + roi (1 + 2 log roi ) D4i = i roi ,
roi

D1i + rri2 D2i + log rri D3i + rri2 log rri D4i = 0 ,
2rri D2i +

1
D3i + rri (1 + 2 log rri ) D4i = i rri .
rri

(2.20a)
(2.20b)

(2.20c)
(2.20d)

These equations are solved to determine D1i , D2i , D3i and D4i . With these coefficients,
the tangential velocity inside the annular cavity is written as

v ( ) ( r ) =

= [2 D2i r + 3i + D4i r + 2 D4i r log r ] .


r
r

(2.21a)

The tangential shear stress on the fluid inside the cavity between adjacent teeth is defined
as

(rfir) (r ) = r

D
D
v ( ) ( r )
[
] = 2r[ 33i 4i ] .
r
r
r
r

(2.21b)

The force acting on the fluid at (r , ) inside the cavity between adjacent teeth becomes

Frfi (r , ) = Ac (rfir) (r ) = 2Ac [

D3i
r2

D4i ] ,

39

(2.22a)

where Ac is the area of the cavity shown in Fig. 2.2(b). The torque loss due to the
swirling flow of the lubricant inside the cavity is then found as

roi

Trfi =

Frfi (r , )dr =2Ac (roi rri )[ roi 3riri D4i ] .

(2.22b)

rri

Finally, the power loss due to root filling inside one tooth cavity can be calculated as the
product of Trfi and i . If there are an average of ni cavities below the oil level where
eddies take place, the power loss of gear i due to root filling is then given as

Prfi = 2ni Ac i (roi rri )(

D3i
D4i ) .
roi rri

(2.23)

The power loss of the gear pair due to root filling becomes Prf = Prf 1 + Prf 2 .

2.5 Oil Pocketing Power Loss


As two adjacent teeth of a gear approach the interface of the mating gear, the
cavity between them is intruded upon by a tooth of the mating gear. This results in a
swift reduction in the volume of the cavity, forcing the lubricant out through the openings
at the ends and the backlash. Figures 2.3(a) to 2.3(d) illustrate the side view of a pair of
gears at a sequence of four discrete rotational positions. Here, a cavity of gear 1 (the
driving gear), reduces in area as it moves to the center of the gear mesh, reaching its
40

minimum in Figure 2.3(d). Similarly, an example cavity of the mating gear is shown to
experience the same continuous contraction before it reaches a minimum at the position
shown in Figure 2.3(b).

In general, there are multiple cavities that are squeezed

simultaneously, exactly the same way, phase-shifted at the base pitch of the gear pair.
This squeezing action of the oil through the respective flow areas results in power loss,
which adds to the total spin loss experienced by the gear pair.
In the meshing zone, control volumes are defined by involute surfaces and root
profiles of the mating gear teeth, with the number of control volumes depending on the
involute contact ratio of the gear pair. Viewing one of the control volumes in threedimension in Figure 2.4, the end flow areas are defined by the tooth height and several
involute and root profile parameters of the respective gears, whereas the backlash flow
area is defined by the shortest chord length that joins the trailing edge of the tooth on the
driving gear and the involute surface of the corresponding mating tooth on the driven
gear. The first task here is to calculate the velocity of the lubricant escaping through the
backlash and end flow areas. This will be done by means of the continuity equation,
assuming an incompressible fluid. Since each control volume varies with rotation of the
gears, it is necessary to determine the flow areas as a function of angular position of the
mating gears. After the velocity of the lubricant squeezed out through the respective flow
areas is calculated, the power loss due to pocketing is determined by using the principle
of conservation of momentum. The above approach does not consider the effects of air
ingestion, which can take place at higher rotational speeds. In such cases, a compressible
fluid formulation is required that is a steep departure from the incompressible flow
formulation presented here.

This problem will be tackled separately in Chapter 3.


41

2
(0)
H12

(0)
H11

(a)

m=0
(0)
H 21

(m )

H12 1

(m )

H11 1

(b) m = m1
(m )

H 21 1

(m )

H11 2

(c) m = m2

(m )

H11 3

(d) m = m3

Figure 2.3 Illustration of a side view of fluid control volumes of the gear mesh interface
at different rotational positions 0 < m1 < m2 < m3 .

42

The initial mesh position for this pocketing formulation is taken as the one when the tip
of a tooth of the driven gear first comes in contact with the involute surface of tooth of
the driving gear at its start of active profile (SAP) as shown in Figure 2.3(a). At this
initial position, a given control volume H ij(0) , the j-th control volume of the i-th gear at
position m = 0 , is associated with the corresponding backlash flow area Ab(0)
,ij and end
(0)
flow Ae(0)
,ij areas. These leakage areas, and subsequently the volume of H ij , change

with rotation of the gears. The first task involves the computation of backlash and end
flow areas at a rotational incremental angle of im = i 0 + mi M ( m [0, M 1] ) where

i 0 = [(rsi2 rbi2 ) 1]1/ 2 and i = [(roi2 rbi2 ) 1]1/ 2 [(rsi2 rbi2 ) 1]1/ 2 . Here, it is noted that
the flow areas of a control volume at m = M are equal to those of the control volume
(M )
(0)
(M )
preceding it at m = 0 , i.e. Ae(0)
,i ( j +1) = Ae,ij and Ab ,i ( j +1) = Ab ,ij .

2.5.1 Backlash Flow Area


Figures 2.5 and 2.6 show the transverse geometry of two spur gears in mesh at an
arbitrary rotational position m . With gear 1 designated as the driving gear, the first step
is to establish the necessary angles and distances at this initial position, focusing on the
(0)
in Figure 2.3(a). The centers of the gears are
first cavity of gear 1 marked as H11

marked as O1 and O2 , with the operating center distance O1O2 = rp1 + rp 2 = e , where

rpi is the pitch radius of gear i.

At this starting position, point B , which is the

intersection of the leading profile of driving gear 1 with the outside (major) circle of gear
2, coincides with the contact point C . This point also represents the SAP of gear 1 so
43

Ae(,mij)
Ae(,mij)

)
Ab( m
,ij

Figure 2.4 Three-dimensional representation of a control volume showing backlash and


end flow areas.

44

that O1B (0) = rs1 , where rs1 is the radius at the SAP for the driving gear. Points D and
E are located at the tip of the leading and trailing profiles of gear 1, whereas point F
marks the intersection of the major circle of the driven gear 2 with the trailing tooth
profile of gear 1. Distances O 1 D = O1E = ro1 remain unchanged with rotational position

m , where ro1 is the outside radius of gear 1. Using basic geometry, the angles to be
defined at m = 0 are given as

1(0) = cos 1[

e2 + rs21 ro22
],
2e rs1

(2.24a)

1
(0)
2 = cos [

e2 + ro22 rs21
],
2e ro 2

(2.24b)

1(0)

(0)
2

=(

ro21
rb21

1)

= cos [

(0) =

1/ 2

1 rb1

cos [

ro1

] (

e2 + (O2 E (0) )2 ro21


2e O2 E (0)

rs21
rb21

1)1/ 2 + cos 1[

rb1 DE
]+
,
rs1
ro1

],

(2.24d)

1(0) 1(0) ,
2

(2.24e)

where DE is the tooth thickness of gear 1 at its tip.


calculated, angles CO2 J (0) = (

(2.24c)

Once the initial angles are

2 DE

) and CO2G (0) = CO2 J (0) JO2G (0) are defined


n2 ro 2

by letting O2 J = ro 2 and assuming more over that O2 E (0) O2G (0) , where

O2 E (0) = (e2 + ro21 2ero1 sin[(0) ])1/ 2 ,


45

(2.24f)

O2

2( m)

2( m)

E
C

1( m)

1( m)

1
(m)
O1

Figure 2.5 Geometry of two gears in mesh at an arbitrary position m .


46

P
M
N
G
E

m)
Ae(,1
j

)
Qb( m
,1 j

C
J

F
B

Qt(,1mj)

Figure 2.6 Definition of the end area at an arbitrary position m .

47

JO2G (0) = (

ro22
rb22

1)1/ 2 cos 1[

rb 2
r
O E (0) 2
] [( 2
) 1]1/ 2 + cos 1[ b 2(0) ] . (2.24g)
ro 2
rb 2
O2 E

(0)
Accordingly, EO2G (0) = CO2 J (0) JO2G (0) (0)
2 and the initial backlash length EG

(the shortest distance between point D on gear 1 and the involute profile of gear 2) is then
given as EG (0) = O2 E (0) EO2G (0) . Thus, the backlash flow area for the first control
(0)
volume H11
at the initial position can then be calculated as

Ab(0)
,11 = b EG

(0)

(2.25)

where b is the effective face width of the gear pair.


With the same gear pair now at an arbitrary rotational position m , the required
angles at that position are given as

( m ) = (0) + (m 1) 1m ,

(2.26a)

1( m)

r
DE ro21
O1C ( m) 2
1/ 2
1 rb1
=
+ ( 2 1) cos [ ] [(
) 1]1/ 2 + cos 1[ b1( m) ] , (2.26b)
ro1
ro1
rb1
rb1
O1C

1( m )

( m) ( m)
2 1 ,

=
( m)
( m)
+ 1 2 ,

( m )

(m)

>
2

48

(2.26c)

(2m ) = sin 1[

O1C ( m ) sin(1( m ) )
O2C ( m )

],

(2.26d)

where distance O2C ( m ) = e2 + (O1C ( m ) )2 2eO1C ( m ) cos[1( m ) ]

is calculated from

known

defines

angles

and

lengths.

One

further

distance

O2 E ( m ) = e2 + ro21 2e ro1 cos[1( m ) + 1( m ) ] and the angle

(2m )

( m)
( m)
1 ro1 sin[1 + 1 ]
= sin [
] (2m ) .
( m)
O2 E

(2.26e)

With O2 J ( m ) = O2 J = ro 2 , CO2G ( m ) = CO2 J ( m ) JO2G ( m ) , assuming O2 E (0) O2G (0) ,


and further calculating

CO2 J ( m) = CO2 J (0) (

JO2G (0) = (

ro22
rb22

ro22
rb22

1) 2 + cos 1[
1

1)1/ 2 cos 1[

1
rb 2
r
O C ( m) 2
] + [( 1
) 1] 2 cos 1[ b 2( m) ] , (2.26f)
ro 2
rb1
O1C

rb 2
r
O E (m) 2
] [( 2
) 1]1/ 2 + cos 1[ b 2( m) ] ,
ro 2
rb 2
O2 E

(2.26g)

the angle EO2G ( m ) = CO2 J ( m ) JO2G ( m ) (2m ) is calculated and the backlash length
EG ( m ) is estimated as

EG ( m ) = O2 E ( m ) EO2G ( m ) . It is noted here that EG ( m ) is

defined by drawing a line from the tip of the trailing edge of the driving gear to the
involute surface of the driven gear along the direction tangent to the major diameter of
49

gear 2. The error associated with this estimation is minimal while the computational
( m)
at
benefits are quite significant. The backlash flow area for the control volume H11

position m is then calculated as

)
( m)
Ab( m
,
,11 = b EG

(2.27)

where b is the effective face width of the gear pair. The same procedure is applied to the
( m)
other control volumes as well. In Figure 2.3, for instance, calculations for H12
are also

)
( m)
( M + m)
repeated the same way to determine Ab( m
, calculations of all
,12 . Since Ab ,i ( j +1) = Ab ,ij

control volumes at M discrete positions covering one base pitch are sufficient to
determine the change of backlash areas of all cavities throughout the gear mesh.
)
Likewise, the same formulation is also applied to the backlash areas Ab( m
,2 j of control

volumes H 2( mj) of gear 2.

2.5.2 End Flow Area


)
Focusing on a control volume H1( m
in Figure 2.6, the end area Ae(,1m)j can be
j

defined as

)
( m)
Ae(,1m)j = Qc 2 Qt(,1mj) Qb( m
,1 j QBCK ,

(2.28)

50

where Qc 2 is the total tooth cavity area bounded by the major diameter and two adjacent
teeth of gear 2 shown in Figure 2.7(a), Qt(,1mj) is the portion of Qc 2 overlapped by the
tooth of gear 1 (area defined by points B, D, E and F) as represented in Figure 2.7(b),
)
Qb( m
,1 j is the excluded area of the segment along the trailing profile (defined by the points
( m)
E, G, J and F) as seen in Figure 2.7(c) and QBCK
is the small portion of the area shown

in Figure 2.6, as bounded by the involute segment BC of gear 1 and CK of gear 2. At


(0)
( m)
the initial position, QBCK
= 0 . As the gears rotate, area QBCK
increases while remaining

typically very small compared to the total area of the cavity Qc 2 . For this reason, it is
neglected here while it can be included, if desired.

2.5.2.1

Calculation of Area Qc 2 : From Fig. 2.7(a), the transverse area Qc 2

between the two adjacent gear teeth can be written as

Qc 2 = QJO2 K QJO2M QKO2 N QMO2 N + QPMN .

(2.29)

Here, QJO2 K is the area of the segment defined by lines O2 J , O2 K (defined from the
center of gear 2 to the tips of two adjacent teeth), area QKO2 N represents the region
between the involute profile segment KN and the segment defined by the angle KO2 N ,
and area QJO2 M is the region between the involute profile segment JM and the segment
defined the angle JO2 M , where M and N are points where the base circle of gear 2

51

O2
(m)
QJO
M

( m)
QPMN

rrf

( m)
QMO
2N
(m)
QKO
2N

ri 2

O2

C
J

F
(m)
QGO
E
1

(m)
QDO
1B

ro 2
52

( m)
QJO
F
1

(m)
QEO
D

(m)
QFO
E
1

(b)

(a)

(c)

O1

Figure 2.7 Parameters used in calculation of (a) the total tooth cavity area Qc 2 ,(b) the overlap area Qt(,1mj) and (c) the excluded
)
area Qb( m
,1 j , all at an arbitrary position m .

52

intersects the involute profiles of gear 2, as shown in Figure 2.7(a). Area QMO2 N is
defined by the segment MN and the angle MO2 N . Finally, the area bounded by points
M, N and P is represented by QPMN where P is a point on the root circle. Assuming that
QPMN is the area of a semi-circle at the root fillet radius, rrf 2 , and by using area
( m)
( m)
integrals, while at the same time recognizing that QJO
M = QKO N for teeth having
2

symmetric involute profiles in both drive and coast sides, area Qc 2 can be written from
Eq. (2.29) as

Qc 2 = 12 ro 2 (

2ro 2
To 2 ) 2 I JM ()drd
N2
JO2 M
,
2rb 2
2
1
1
2 rb 2 (
Tb 2 ) + 2 rrf 2
N2

(2.30)

where radius I JM () represents the radius of a point along the involute profile at a given
angle , To 2 and Tb 2 are circular tooth thicknesses of gear 2 at the tip and base circle,
and N 2 is the number of teeth in gear 2. A simplified version I JM () = rb 2 [1 + (3) 2 / 3 ]
is used here by expressing the involute profile equation in the Taylor series form and
retaining the first order terms [2.12]. With this, the area integral bounded by the involute
profile segment JM is given by

JO2 M

2.33
I JM ()drd = 12 rb22 [(M J ) + 0.46(2.33
M J )

+ 1.25(1.67
M

1.67
J )]
53

(2.31)

where, J and M denote the angles measured from the horizontal line X 2O2 to the
lines O2 J and O2 M , respectively. The value of area Qc 2 is obtained in closed-form by
substituting Eq. (2.31) into Eq. (2.30).

2.5.2.2 Calculation of Area Qt(,1mj) : In order to calculate the overlap area Qt(,1mj)
)
associated with the control volume H1( m
j shown in Figure 2.6, the intersection points B

and F of the outside circle of the driven gear with the involute surfaces of the driving gear
must be defined.

For this purpose, distances O1F ( m ) and O1B ( m ) are determined

numerically such that O2 B ( m ) = O2 F ( m ) = ro 2 in Figure 2.7(b). The overlap area Qt(,1mj) is


then given as

(m)
(m)
(m )
Qt(,1mj) = QDO
B + QEO1D + QFO E QFO B .
1

(2.32)

( m)
Here, QDO
B is the area of the region bounded by the involute profile segment BD , lines
1

(m)
O1B , O1D and angle BO1D , area QFO
is the area of the region bounded by the
1E
(m)
involute profile segment FE , lines O1F , O1E and angle FO1E , area QFO
is the area
1B

of the region bounded by the arc BF , lines O1B , O1F and angle BO1F and, finally,
area QEO1D is the area of the region bounded by the lines ED , O1D , O1E and the

54

angle DO1E . With this nomenclature, Figure 2.7(b) shows all of the four areas used in
Eq. (2.32) to define Qt(,1mj) . Equation (2.32) can then be written in integral form as

Qt(,1mj) =

1
I BD ()drd + ro1To1 + I FE ()drd I BF ()drd .
2
DO B
FO E
FO B

(2.33)

Here, functions I BD () and I FE () define the involute profiles on the drive and coast
sides, both approximated by I () = rb1[1 + 0.5(3)2 / 3 ] , so that their respective integral
terms can be evaluated as

BO1D

FO1E

I BD ()drd = 12 rb21[( B D ) + 0.46{( R D )2.33 (R B ) 2.33}


1.67

+ 1.25{( R D )

1.67

( R B )

}]

I FE ()drd = 12 rb21[( E F ) + 0.46{( E Q )2.33 ( F Q ) 2.33}


1.67

+ 1.25{( E Q )

1.67

( F Q )

, (2.34a)

, (2.34b)

}]

where, angles B , D , E and F are measured from a horizontal line X1O1 , which is
perpendicular to the lines joining the gear centers, to the lines O1B , O1D , O1E and O1F ,
(m)
respectively, as shown in Figure 2.7(b). For evaluating the area QFO
B , the radius r of
1

any point on the arc BF can be expressed in terms of the gear angle as
I BF () = e sin ro22 e2 sin 2 . With this, one obtains
55

I BF ()drd = 12 ro22 ( B F ) 14 e2 [sin(2 B ) sin(2 F )]

BO1F

+ 12 ro22 [sin 1 (

e cos B
e cos F
) sin 1 (
)]
ro 2
ro 2

+ 12 ero 2 [cos B 1 (

(2.34c)

e cos B 2
e cos F 2
) cos F 1 (
) ],
ro 2
ro 2

with angles B and F defined in Figure 2.7(b). Area Qt(,1mj) is then determined by
substituting Eqs. (2.34a) - (2.34e) into Eq. (2.33).

2.5.2.3

)
Calculation of Area Qb( m
,1 j :

)
Next, the excluded area Qb( m
,1 j of the

segment along the trailing profile (defined by the points E, G, J and F in Figure 2.6) of
)
the control volume H1( m
j is written as

)
( m)
( m)
( m)
( m)
Qb( m
,1 j = QGO E + QJO G QJO F QFO E ,
1

(2.35)

(m)
where any area QXO
Y is defined as the area of the segment between the line XY and the
1

angle XO1Y , as shown in Figure 2.7(c). Expressing Eq. (2.35) in terms of the area
integrals, one writes

)
( m)
Qb( m
,1 j = QGO E +
1

JO1G

I JG ()drd

JO1F

56

I JF ()drd

FO1E

I FE ()drd ,

(2.36)

where functions I JG () and I FE () define the involute profiles on the drive side of gear
2 and the coast side of gear 1, respectively, whereas function I JF () defines the segment
JF that is a portion of the outside circle of gear 2 intersecting gear 1 profile at points F
( m)
and B , respectively. Area QGO
E is approximated by a triangle with known dimensions
1

for its edges EG ( m) , O1E ( m) and O1G ( m) . It is calculated as

( m)
( m)
QGO
)( s O1E ( m) )( s O1G ( m ) ) ,
E = s ( s EG
1

(2.37a)

where s = 12 ( EG ( m ) + O1E ( m ) + O1G ( m ) ) . Given I JF () = e sin ro22 e2 sin 2 , area


( m)
QJO
F is calculated as
1

I JF ()drd = 12 ro22 ( F J ) 14 e 2 [sin(2 F ) sin(2 J )]

JO1F

+ 12 ro22 [sin 1 (

e cos J
e cos F
) sin 1 (
)]
ro 2
ro 2

+ 12 ero 2 [cos F 1 (

(2.37b)

e cos J 2
e cos F 2
) cos J 1 (
) ],
ro 2
ro 2

where angles F and J are defined from the horizontal line X1O1 to the lines O1F and
(m)
O1 J , respectively. Area QFO
E is calculated from Eq. (2.34b), with angles E and F ,
1

defined from the horizontal line X1O1 to the lines O1E and O1F , respectively, taken as

57

the upper and lower limits of integration, while the area

( m)
QJO
G
1

with

2
I JG () = e sin rinv
e 2 sin 2 is obtained as

2
I JG ()drd = 12 rinv
(G J ) 14 e2 [sin(2G ) sin(2 J )]

JO1G
2
+ 12 rinv
[sin 1 (

e cos G
e cos J
) sin 1 (
)]
ro 2
ro 2

+ 12 erinv [cos G 1 (

(2.37c)

e cos G 2
e cos J 2
) cos J 1 (
) ],
ro 2
ro 2

where rinv is a length that varies as a function of the rotational angle from O1 J to O1G .
Substituting Eqs (2.37a) to (2.37c) into Eq. (2.36) will give a closed form expression for
)
excluded area Qb( m
,1 j .

2.5.3 Power Loss due to Oil Pocketing


)
( m)
With Ae(,mij) (one at each side) and Ab( m
at
,ij defined, the volume of the cavity H ij

a rotational position m is given by Vij( m) = bAe(,mij) . Next, the velocity of the oil squeezed
out through the backlash and end flow areas are determined by using the integral form of
the continuity equation [2.17]

d
dV = v n dS ,
dt V
S

(2.38a)

58

where V is an arbitrary volume, is the density of the oil, v is the velocity vector and
n is the outward unit vector normal to the surface element S of the arbitrary volume V.

Integrating the above equation and assuming one-dimensional flow, the general flow
velocity escaping through a discharge area can be written in terms of a flow area A ,
rotational speed and the rate of change of flow volume as

v=

1 dV
1 dV
=
.
A dt
A d

(2.38b)

The negative sign here vanishes since dV d is negative during pocketing, as each
successive volume is smaller than the previous, until the minimum position is reached.
Thus, for the first control volume, the velocity of the lubricant escaping through the
backlash flow area and end flow area can be written, respectively, as

)
vb( m
,ij =

(m)
( m) dVij
i ,
) d
Ab( m
,ij

(2.39a)

( m)

ve( ,mij)

(1 ( m) ) dVij
=
i ,
d
2 Ae(,mij)

(2.39b)

)
( m)
( m)
where ( m) = Ab( m
,ij (2 Ae,ij + Ab,ij ) is a flow factor defined as ratio of the backlash flow

area to the total flow area. Since there are two areas of end leakage and also one area of
backlash leakage, a three-dimensional analysis would be required in calculating the flow
59

velocity along each flow area. A portion of the oil is squeezed out through the end
leakage area while the rest of it is discharged through the backlash leakage area. This
flow factor ( m ) is introduced in order to ease the computational demand associated with
such a three-dimensional computation, which is beyond the scope of this dissertation.
Here, including ( m ) in Eq. (2.39a, b), in essence, captures the portion of lubricant
escaping through the backlash and end leakage areas, respectively, due to the pumping
action of the gears. For a gear pair with a very large face width relative to its outside
diameter, ( m) 1 , as the backlash flow area dominates over the end leakage area, while
the end leakage area dominates over the backlash flow area for gears with narrow face
width, and hence, ( m) 0 .

The pressure of the control volume H ij( m) at the backlash and end flow areas can be
estimated by using Bernoullis principle [2.17], respectively, as

)
( m 1) 1
1) 2
( m) 2
pb( m
+ 2 [(vb( m
,ij = pb ,ij
,ij ) (vb,ij ) ] ,

(2.40a)

)
( m 1) 1
1) 2
( m) 2
pe( m
+ 2 [(ve( m
,ij = pe,ij
,ij ) (ve,ij ) ] .

(2.40b)

The conservation of momentum principle [2.17] gives the forces acting on the oil
escaping through the backlash and end areas of H ij( m) as

60

)
( m)
( m) ( m)
Fb(,mij ) = pb( m
,ij dAb ,ij = pb ,ij Ab ,ij ,

(2.41a)

S
)
( m)
( m)
( m) ( m)
Fe(,m
ij = pe,ij dAe,ij = pe,ij Ae,ij .

(2.41b)

The power loss due to pocketing at H ij( m) is equal to the product of the discharge
velocity and the pocketing force:

)
( m) (m )
( m) ( m)
Pp( m
,ij = vb ,ij Fb ,ij + 2ve,ij Fe,ij .

(2.42a)

This represents only the power loss for the j-th control volume at the m-th rotational
position for gear i. The same procedure is applied to all J1( m ) control volumes of gear 1
and J 2( m ) control volumes of gear 2 to determine the total instantaneous power loss of a
gear pair due to pocketing that is then averaged over the total number of rotational
increments, M, to find the average power loss as

1
Pp =
M

J 2( m )
J1( m )

m
(
)
( m)

Pp,1 j + Pp,2 j .
m =1
j =1
j =1

(2.42b)

Additional power losses can take place in the meshing zone due to secondary
effects loss such as lift-off of the lubricant and acceleration of the lubricant trapped in the
61

meshing zone in the circumferential direction. Such effects were not included in this
paper since they were found to be negligibly small when compared to the Pp values
predicted above.

2.6 Example Oil Churning Analysis

The drag and pocketing components of the total gear pair spin loss PT defined in
Eq. (2.2) are determined by using Eqs. (2.7), (2.13) (or Eq. (2.17) if the regime is
turbulent), (2.23), and (2.42b), respectively. These equations, all given in closed-form,
provide the total spin loss as a function of gear parameters, lubricant properties, and
operating conditions.
In order to exercise the proposed gear pair spin loss formulation, two unity-ratio
spur gear pairs having different modules are considered: a 40-teeth pair at a module
m% = 2.32 mm and a 23-teeth gear pair having m% = 3.95 mm. The 40-teeth gear pair is
considered to have three face width values of b = 14.7, 19.5 and 26.7 mm, whereas the
23-teeth gear pair has a facewidth of b = 19.5 mm. The gears are operated at a center
distance of 91.5 mm, under no load (i.e. zero transmitted torque).

Also, since the

example analysis is concerned with computing power losses involved in the case of
unloaded conditions, the surface roughness of the gears does not play any role in the
subsequent analyses. Table 2.1 lists all essential parameters of these two example gear
pairs. In addition to b and m% , rotational speed (in rpm), oil temperature T (in C), and
the oil immersion depth parameter h are also varied to quantify their influence on the
62

_______________________________________________________________
Parameter

23T gears

40T gear

_______________________________________________________________
Number of teeth

23

40

Module, mm

3.95

2.32

Pressure angle, deg

25.0

28.0

Base diameter, mm

82.34

81.89

Outside diameter, mm

100.34

95.95

Circular tooth thickness, mm

6.435

2.925

Root fillet radius, mm

1.63

0.83

Face width, mm

19.5

14.7, 19.5, 26.7

Center Distance, mm

91.5

91.5

_______________________________________________________________

Table 2.1 The design parameters of the example spur gears.

63

resultant PT . The rotational speed ranged from = 1000 rpm to = 4000 rpm , with
oil immersion level ranging from h = 0.5 , which refers to an oil immersion level of 25 %
of the pitch radius of the gear to h = 1.5 , which refers to an oil immersion level of 75 %
of the pitch radius of the gear. Also, the operating oil temperature ranged from 20 C to
100 C. It is to be noted that under the oil immersion depths considered in the example
analysis, the lubricant properties are that of the oil itself with air not taking any part in the
formulations.
It is useful to mention here that each analysis shown in the following figures took
less than a second of CPU time on a personal computer, indicating that the proposed
model is indeed suitable as a practical design tool.
In Figure 2.8(a), the influence of T on PT is demonstrated for m% = 2.32 mm,
b = 19.5 mm and h = 1.0 . Here, a transmission fluid ( = 0.217, 0.035 and 0.011 Pa-s
and = 847.3, 823.3 and 799.3 kg / m3 at T = 20, 60 and100 C, respectively) is
considered as the lubricant. As shown in Figure 2.8(a), PT increases with reduction in T
(increased ) regardless of the value of . For example, spin loss at T = 20 C and
= 4, 000 rpm is about 0.3kW higher than PT value at the same and at T = 100 C.
This difference is due to significant changes in viscosity of the oil (about 15 times higher
viscosity at 20 C than at 100 C, which changes Pdpi and Prfi linearly (Eqs. (2.7) and
(2.23)), while it has a more complex influence on Pdfi and Pp . This predicted effect of
viscosity conflicts with the experimental observations of Luke and Olver [2.8] and that of
64

(a)

1.0

(b)

20 o C

h = 1.5

0.8

60 C

h = 1.0

100 o C

h = 0.5

0.6
PT
[kW]

0.4

0.2

0.0
(c)

1.0

0.8

(d)

b = 26.7 mm

% = 3.95 mm
m

b = 19.5 mm

m% = 2.32 mm

b = 14.7 mm

0.6
PT
[kW]

0.4

0.2

0.0
0

1000

2000

3000

4000 0

1000

2000

3000

4000

Speed [rpm]

Figure 2.8 Effect of (a) temperature T , (b) oil level parameter h , (c) face width b and
(d) gear module m% on total spin power loss PT .

65

Changenet and Velex [2.14] while agreeing well with experiments described in Chapter
4, which talks about the validation of the oil churning and windage power loss models.
In Figure 2.8(b), the influence of the quasi-static oil levels on PT is demonstrated for a
gear pair having m% = 2.32 mm and b = 19.5 mm, operated in lubricant at a temperature of
80 C ( = 0.0185 Pa-s and = 811.3 kg / m3 ). Here, the oil levels h = 0.5 , 1.0 and 1.5
represent 25, 50 and 75% immersion of the gears in oil, according to Figure 2.1. As all
components of PT are dependent on h , significant increases in PT are observed with
increased immersion depth, with almost 2.5 times larger PT values at h = 1.5 compared
to those at h = 0.5 .
Figure 2.8(c) shows the influence of b on PT for a gear pair having m% = 2.32 mm
operating at h = 1.0 and 80 C. Here, the gear pair having b = 26.7 mm experiences
about 2.8 times higher PT than the gear pair with b = 14.7 mm. The components of PT
for the gear pair with m% = 2.32 mm and varying face width values are shown in Figure
2.9. Here, in line with the formulation given above, the major changes with b take place
in Pdp and Pp components of the power loss as shown in Figures 2.9(a) and 2.9(d).
More importantly, Figure 2.9 indicates that Pp and Pdf are the two most significant
components of oil churning power loss of this gear pair while contributions of Pdp and

Prf to PT are both quite negligible. For instance, the power loss values of the gear pair
having b = 19.5 mm at = 4, 000 rpm are such that Pp is about 74.4% of PT followed
by 25.1% for Pdf , while the other two components Pdp and Prf are 0.48% and 0.02%
66

(a) Pdp

0.0030

(b) Pdf

0.20

b = 26.7 mm
b = 19.5 mm

0.0024

0.15

Power Loss [kW]

b = 14.7 mm

0.0018

0.10
0.0012

0.05

0.0006

0.00

0.0000

(c) Prf

0.00015

0.00012
Power Loss [kW]

(d) Pp

0.8

0.6

0.00009
0.4
0.00006
0.2
0.00003

0.0

0.00000
0

1000

2000

3000

4000

1000

2000

3000

4000

Speed [rpm]

Figure 2.9 Components of PT for a gear pair having m% = 2.32 mm and b = 26.7, 19.5

and 14.7 mm at h = 1.0 and 80 C; (a) Pdp , (b) Pdf , (c) Prf , and (d) Pp .

67

of PT , respectively.

The last comparison in Figure 2.8(d) illustrates the influence of module m% on PT


for b = 19.5 mm, h = 1.0 and oil at an operating temperature of 80 C. Here, the coarser
pitch gear pair with m% = 3.95 mm is predicted to have more spin power loss than the gear
pair having m% = 2.32 mm. For instance, at = 4, 000 rpm, PT for m% = 3.95 mm is about
0.1 kW higher than that of the gear pair having m% = 2.32 mm.

As Pdf and Pdp

components are not expected to be influenced significantly by tooth size according to the
formulations presented, and Prf is rather negligible, this difference in PT values with
module can be attributed primarily to the pocketing losses.
As one last observation from Figures 2.8 and 2.9, the value of PT changes with
( = 2 60 ) in a non-linear fashion. For instance, in Figure 2.8(c) for b = 19.5
mm, PT 2.11 while the corresponding components of the power loss in Figure 2.9
reveal the dependency of the individual components of power loss on speed as

Pdf 2.5 , Pdp 2 , Prf and Pp 1.8 .

2.7 Concluding Remarks

A physics-based fluid mechanics model was proposed in this chapter to predict


power losses of gear pairs due to oil churning. Individual formulations were proposed for
evaluating drag, root filling and pumping components of the gear pair spin losses by
treating power losses associated with pumping of the oil from the gear mesh, drag forces
68

induced by the oil on a rotating gear body along its periphery and faces, and finally,
eddies formed in the cavities between adjacent teeth. The model was applied to a unityratio spur gear pair to quantify the individual contributions of each component to the total
oil churning power loss. At the end, the influence of operating conditions, selected gear
geometry parameters, module and face width, and lubricant properties on oil churning
power losses were quantified for the example gear pair.

69

References for Chapter 2

[2.1]

Xu, H., Kahraman, A., Anderson, N.E., Maddock, D.G., 2007, Prediction of
Mechanical Efficiency of Parallel-Axis Gear pairs, ASME Journal of Mechanical
Design, 129, 58-68.

[2.2]

Martin, K. F., 1978, A Review of Friction Predictions in Gear Teeth, Wear, 49,
201238.

[2.3]

Daily, J. W. and Nece, R. E., 1960, Chamber Dimensional Effects on Induced


Flow and Frictional Resistance of Enclosed Rotating Disks, ASME Journal of
Basic Engineering, 82, 217232.

[2.4]

Mann, R.W. and Marston, C.H., 1961, Friction Drag on Bladed disks in
Housings as a Function of Reynolds Number, Axial and Radial Clearance and
Blade Aspect Ratio and Solidity, ASME Journal of Basic Engineering, 83 (4),
719-723.

[2.5]

Bones, R.J., 1989, Churning Losses of Discs and Gears Running Partially
Submerged in Oil, Proceedings of the ASME 5th International Power
Transmission and Gearing Conference, Chicago, 355-359.

[2.6]

Von Karman, T., 1921, On Laminar and Turbulent Friction, Z. Angew. Math.
Mech., 1, 235-236.

70

[2.7]

Terekhov, A.S., 1975, Basic Problems of Heat Calculation of Gear Reducers,


JSME International Conference on Motion and Powertransmissions, Nov. 23-26,
1991, 490-495.

[2.8]

Luke, P., and Olver, A., 1999, A Study of Churning Losses in Dip-Lubricated
Spur Gears, Proceedings of the Institute of Mechanical Engineers, Part G,
Journal of Aerospace Engineering, 213, 337346.

[2.9]

Ariura, Y., Ueno, T., and Sunamoto, S., 1973, The lubricant churning loss in
spur gear systems, Bulletin of the JSME, 16, 881-890.

[2.10] Akin, L. S., and Mross, J. J., 1975, Theory for the Effect of Windage on the
Lubricant Flow in the Tooth Spaces of Spur Gears, ASME Journal of
Engineering for Industry, 97, 12661273.
[2.11] Akin, L. S., Townsend, J. P., and Mross, J. J., 1975, Study of lubricant jet flow
phenomenon in spur gears, Journal of Lubrication Technology, 97, 288-295.
[2.12] Pechersky, M. J. and Wittbrodt, M. J., 1989, An analysis of fluid flow between
meshing spur gear teeth, Proceedings of the ASME 5th International Power
Transmission and Gearing Conference, Chicago, 335342.
[2.13] Diab, Y., Ville, F., Houjoh, H., Sainsot, P., and Velex, P., 2005, Experimental
and Numerical Investigations on the Air-Pumping Phenomenon in High-Speed
Spur and Helical Gears, Proceedings of the Institution of Mechanical Engineers,
Part C: J. Mechanical Engineering Science, 219, 785-800.
[2.14] Changenet, C., and Velex, P., 2007, A Model for the Prediction of Churning
Losses in Geared Transmissions Preliminary Results, ASME Journal of
Mechanical Design, 129(1), pp. 128-133.
71

[2.15] Hhn, B.R., Michaelis, K., Vllmer, T., 1996, Thermal Rating of Gear Drives:
Balance between Power Loss and Heat Dissipation, AGMA, Fall Technical
Meeting, 96FTM8, pp. 1-12.
[2.16] Changenet, C., and Velex, P., 2008, Housing Influences on Churning Losses in
Geared Transmissions, ASME Journal of Mechanical Design, 130 (6), 062603, 6
pages.
[2.17] Streeter, L. V., and Wylie, E. B., 1985, Fluid Mechanics, Eighth edition,
McGraw-Hill, New York.
[2.18] Schlichting, H., 1955, Boundary-Layer Theory, McGraw-Hill Book Company,
New York.
[2.19] Petry-Johnson, T. T., Kahraman, A., Anderson, N.E., and Chase, D. R., 2008, An
Experimental Investigation of Spur Gear Efficiency, ASME Journal of
Mechanical Design, 130 (6), 062601, 10 pages.
[2.20] Xu, H., Kahraman, A., Anderson, N.E., Maddock, D.G., 2007, Prediction of
Mechanical Efficiency of Parallel-Axis Gear pairs, ASME Journal of Mechanical
Design, 129, 58-68.
[2.21] Martin, K. F., 1978, A Review of Friction Predictions in Gear Teeth, Wear, 49,
201238.
[2.22] Daily, J. W. and Nece, R. E., 1960, Chamber Dimensional Effects on Induced
Flow and Frictional Resistance of Enclosed Rotating Disks, ASME Journal of
Basic Engineering, 82, 217232.
[2.23] Mann, R.W. and Marston, C.H., 1961, Friction Drag on Bladed disks in
Housings as a Function of Reynolds Number, Axial and Radial Clearance and
72

Blade Aspect Ratio and Solidity, ASME Journal of Basic Engineering, 83 (4),
719-723.
[2.24] Bones, R.J., 1989, Churning Losses of Discs and Gears Running Partially
Submerged in Oil, Proceedings of the ASME 5th International Power
Transmission and Gearing Conference, Chicago, 355-359.
[2.25] Von Karman, T., 1921, On Laminar and Turbulent Friction, Z. Angew. Math.
Mech., 1, 235-236.
[2.26] Terekhov, A.S., 1975, Basic Problems of Heat Calculation of Gear Reducers,
JSME International Conference on Motion and Powertransmissions, Nov. 23-26,
1991, 490-495.

[2.27] Luke, P., and Olver, A., 1999, A Study of Churning Losses in Dip-Lubricated
Spur Gears, Proceedings of the Institute of Mechanical Engineers, Part G,
Journal of Aerospace Engineering, 213, 337346.
[2.28] Ariura, Y., Ueno, T., and Sunamoto, S., 1973, The lubricant churning loss in
spur gear systems, Bulletin of the JSME, 16, 881-890.
[2.29] Akin, L. S., and Mross, J. J., 1975, Theory for the Effect of Windage on the
Lubricant Flow in the Tooth Spaces of Spur Gears, ASME Journal of
Engineering for Industry, 97, 12661273.
[2.30] Akin, L. S., Townsend, J. P., and Mross, J. J., 1975, Study of lubricant jet flow
phenomenon in spur gears, Journal of Lubrication Technology, 97, 288-295.
[2.31] Pechersky, M. J. and Wittbrodt, M. J., 1989, An analysis of fluid flow between
meshing spur gear teeth, Proceedings of the ASME 5th International Power
Transmission and Gearing Conference, Chicago, 335342.
73

[2.32] Diab, Y., Ville, F., Houjoh, H., Sainsot, P., and Velex, P., 2005, Experimental
and Numerical Investigations on the Air-Pumping Phenomenon in High-Speed
Spur and Helical Gears, Proceedings of the Institution of Mechanical Engineers,
Part C: J. Mechanical Engineering Science, 219, 785-800.
[2.33] Changenet, C., and Velex, P., 2007, A Model for the Prediction of Churning
Losses in Geared Transmissions Preliminary Results, ASME Journal of
Mechanical Design, 129(1), pp. 128-133.
[2.34] Hhn, B.R., Michaelis, K., Vllmer, T., 1996, Thermal Rating of Gear Drives:
Balance between Power Loss and Heat Dissipation, AGMA, Fall Technical
Meeting, 96FTM8, pp. 1-12.
[2.35] Changenet, C., and Velex, P., 2008, Housing Influences on Churning Losses in
Geared Transmissions, ASME Journal of Mechanical Design, 130 (6), 062603, 6
pages.
[2.36] Streeter, L. V., and Wylie, E. B., 1985, Fluid Mechanics, Eighth edition,
McGraw-Hill, New York.
[2.37] Schlichting, H., 1955, Boundary-Layer Theory, McGraw-Hill Book Company,
New York.
[2.38] Petry-Johnson, T. T., Kahraman, A., Anderson, N.E., and Chase, D. R., 2008, An
Experimental Investigation of Spur Gear Efficiency, ASME Journal of
Mechanical Design, 130 (6), 062601, 10 pages.

74

CHAPTER 3

WINDAGE POWER LOSSES OF A GEAR PAIR: MODEL FORMULATION

3.1 Introduction
In Chapter 2, a fluid mechanics based oil churning power loss model for a spur
gear pair was proposed. As depicted in Figure 2.1, the spur gear pair was immersed in a
liquid medium (lubricating oil bath). In such a case of gears operating while immersed in
an oil bath, with this method of lubrication often called dip lubrication, the oil churning
losses under dip-lubrication conditions were represented by the sum of (i) the losses due
to oil drag on the faces and periphery of the gears, (ii) the losses associated with eddies
formed during the filling of lubricant in the cavities between adjacent gear teeth and (iii)
losses caused by the displacement of the oil trapped within the gear mesh due to
squeezing and pumping. Examples of geared systems operating under dip-lubrication
condition are ample in low to medium speed automotive transmissions and in industrial
gearbox applications.
When the operating gear speeds are relatively high, the jet-lubrication method is
preferred instead of dip-lubrication, as is the case with most aerospace powertrain and

75

turbine applications. In this method of lubrication, a high-pressure jet of oil is directed


towards the gear mesh interface via a nozzle, and oil is not allowed to accumulate within
the gearbox to cause any oil churning losses. Hence, power losses under jet-lubrication
conditions can mostly be attributed to windage losses due to the interaction between air
or air-oil mixture and the rotating gears. Pocketing power loss due to squeezing of the
compressible air/air-oil mixture from the meshing zone forms the main component of
windage power loss. In addition to pocketing loss, windage drag power loss due to air
drag along the periphery and surfaces of the rotating gears can also be expected to
contribute additionally to the total windage power loss. It was shown experimentally by
Petry-Johnson et al. [3.1] that power losses in dip-lubricated gearboxes with very low
static oil levels can also be dominated by air windage. With the assumption that the oil
jet directed at the gear mesh has a negligible influence, the windage power loss Pw of a
gear pair is given as

Pw = Pwp + Pwd ,

(3.1)

where Pwp is the air pumping/pocketing power loss and Pwd = Pwd1 + Pwd 2 is the total
drag power loss of a gear pair. Meanwhile, the drag power loss of each gear i ( i = 1, 2 ) is
given as the sum of drag losses at its periphery, Pwdpi , and its faces (sides), Pwdfi , as

Pwdi = Pwdpi + Pwdfi .

(3.2)

76

Most of the investigations on gear windage power losses have been experimental
in nature, focusing on losses of gears or disks of various sizes rotating in free air and
surrounded by different enclosures. In one such study, Dawson [3.2] performed a large
number of gear windage experiments using spur and helical gears rotating in free air and
derived empirical formulae to describe windage power losses as a function of gear
geometry, lubricant (air) and speed parameters. Diab et al [3.3] carried out windage
experiments on disks and gears of an assortment of shapes and sizes, and subsequently
came up with quasi-analytical predictions based on fluid flow around a rotating gear and
also through dimensional analysis, to characterize windage loss in pinion-gear pairs.
Their model does not consider the presence of the mating gear and hence, the major
component of windage power loss in the form of fluid compression and expansion taking
place in the meshing zone between gear teeth is not considered in the model. Eastwick
and Johnson [3.4] have recently provided an extensive review of studies on gear windage
to conclude that a physics-based, generalized solution for reducing power loss due to
windage has not yet been well established. Wild et al [3.5] studied the flow between a
rotating cylinder and a fixed enclosure by using a computational fluid dynamics (CFD)
model and accompanying experiments. Similarly, Al-Shibl et al [3.6] proposed a CFD
model of windage power loss from an enclosed spur gear pair.
Review of these studies indicates that there is a need for an accurate and
computationally efficient windage power loss model for predicting power losses of jetlubricated, high-speed spur gear pairs. Accordingly, the main objective of this chapter is
to develop a fluid mechanics-based model for predicting windage losses of spur gear
pairs. This model will consist of a compressible fluid flow gear mesh pocketing power
77

loss model, and an air drag formulation through a modification of the drag power loss
formulations from the oil churning model presented in Chapter 2.

Following the

formulation of the windage power loss model, the validation of the model with respect to
actual windage loss experiments [3.1] and empirical results [3.2] will ensue in Chapter 4.
At the end of this chapter, the model will be exercised for a pair of example spur gear sets
to quantify the contributions of the major components of windage power loss and to
highlight the computational ease of use of the proposed windage power loss model
through a parametric study. In the subsequent sections, formulations will be proposed to
compute Pwp , Pwdpi and Pwdfi , so that the total windage power loss of a spur gear pair
can be determined according to Eq. (3.1).

3.2 Windage Pocketing Power Loss Model


Formulation of the model for power loss due to pumping of air or air-oil mixture
from the gear mesh is identical to that of oil pumping power loss proposed in Chapter 2,
in terms of the changes in discharge areas and control volumes formed between the
mating teeth of the gears as they rotate. The main difference between the windage
pocketing power loss model and the oil churning pocketing model is that the medium
considered here is air or a mixture of air and oil, depending on the aeration ratio, and
hence, the assumption made in Chapter 2 that the fluid is incompressible is no longer
valid: variations in density as the pocket volume changes with rotational position has to
be considered in case of compressible flow, and corresponding variations in pressure,
temperature and velocity of the fluid ejected through the respective discharge areas will
78

follow. The windage pocketing phenomenon is defined through the squeezing action of
any control volume formed between a gear tooth and the mating tooth cavity over
successive rotational positions, as illustrated in Figure 2.3. Looking at Figure 2.3, it is
necessary to conceive the pockets to be filled by air or a mixture of air and oil, instead of
the incompressible lubricant. As the air or air-oil mixture gets squeezed out from the
openings at the ends and at the backlash (as seen in Figure 2.4) due to successive
compression of the control volume till the volume is minimized at a certain rotational
gear position, a pressure differential is created between the ambient atmosphere outside
the control volume and the air-oil mixture or air inside the control volume. In the
process, the mixture is squeezed out at typically high velocities through the discharge
areas. The high velocity of compressible fluid ejected through the flow areas will result
in power losses.
The formulation defined below for the windage pocketing power loss in terms of
compressible fluid flow will first evaluate the discharge areas formed between the
involute profiles of the mating gears. This evaluation procedure for the discharge areas
follows very much on the lines for computing the flow areas in case of the meshing zone
filled with lubricant as laid out in Chapter 2. Next, a methodology to compute the
velocity of the air or air-oil mixture being squeezed out through the end and backlash
flow areas will be proposed through means of the conservation of mass and conservation
of energy principles. The pressure inside the control volume will then be computed using
two different methodologies, which depart from each other at higher operating speeds of
rotation (Mach numbers higher than 0.3): (i) the compressible Bernoulli formulation and,
(ii) isentropic relationship between pressure and density inside control volume. The
79

conservation of momentum principle will be used to compute the force acting on the
atmosphere inside the control volume through the product of the mass flow rate and the
velocity of the fluid escaping through the discharge areas.

Finally, the windage

pocketing power loss will be then be defined as the product of the force acting on the
fluid inside the control volume and the velocity of the compressible fluid being ejected
out of the control volume.
It has to be noted that computation of velocity of the fluid escaping through the
backlash discharge area can be handled in two different ways. One line of thought is to
treat the pockets (control volumes) formed between the mating gear teeth as separate
entities, with no pressure differential being communicated between opposing pockets or
(0)
, which is the first cavity of gear 2 and
control volumes (an example labeled as H 21
(0)
H12
, which is the second cavity of gear 1, both at the beginning of contact, seen in

Figure 2.3(a)). Another way of approach is to consider the pressure differential between
the opposing control volumes, as a result of which, not all of the fluid lying in the
backlash discharge region gets ejected through the backlash discharge area. Some of the
fluid will in fact be transported inside the opposing control volume having lesser pressure
and forcefully ejected through the end discharge area corresponding to that control
volume, reinforcing the velocity of the compressible mixture already being ejected
through those particular end flow areas, by augmenting the mass flow rate of the fluid
ejected through the end area.

In order to alleviate concerns about the physics of

broaching the problem through either means of thought, in this work, the control volumes

80

or pockets will be considered as independent entities, with power loss taking place
through both the discharge areas.
Referring to Figure 2.4, areas of the openings of a given cavity H ij( m) (j-th cavity
)
of gear i) at a rotational position m along its backlash and end (side) openings, Ab( m
,ij and

Ae(,mij) , are defined as in Chapter 2, respectively, as

)
( m)
Ab( m
,
,ij = b EG

(3.3)

)
( m)
Ae(,mij) = Qc 2 Qt(,mij ) Qb( m
,ij QBCK ,

(3.4)

where b is the effective face width of the gear pair and EG ( m) is the backlash length (the
shortest distance between a point on the coast side of gear 1 and the involute profile of
gear 2), seen in Figure 2.6. As defined in Figure 2.7(a), Qc 2 is the total tooth cavity area
bounded by the major diameter and two adjacent teeth of gear 2, Qt(,mij ) is the portion of
)
Qc 2 overlapped by the tooth of gear I, as shown in Figure 2.7(b), Qb( m
,ij is the excluded
( m)
area of the segment along the trailing profile, depicted in Figure 2.7(c), and QBCK
is the

small portion of the area bounded by the involute segment BC of gear 1 and CK of gear
2.
)
( m)
Closed form expressions for Ab( m
,ij and Ae,ij were derived in Chapter 2 based on

gear tooth geometry and gear kinematics. The very same expressions will be used here as
81

well, with the only different that the medium inside the pockets is air or a mixture of air
and oil, and hence, the assumption made in Chapter 2 that the fluid is incompressible is
no longer valid. Meanwhile, as a considerable departure from the incompressible fluid
flow formulation of the earlier oil pocketing model of Chapter 2, the windage pocketing
formulations must use a compressible flow approach to compute the flow velocities
through the end and backlash flow areas, as reiterated in preceding paragraphs. Applying
the principle of conservation of mass [3.7] to an arbitrary control volume, the total rate of
change of mass in a control volume is equal to the sum of the rate of change of mass
accumulated inside the control volume and the total mass exiting (or entering) the control
volume such that

d
dV + v n dS = 0 .
dt V
S

(3.5)

Here V is an arbitrary volume, is the density of the squeezed air (or air-oil mixture), v
is the velocity vector of the air/mixture escaping through the control volume, and n is the
outward unit vector normal to the surface element S of the arbitrary volume V.
Applying Eq. (3.5) to sample control volumes, H ij( m) and H ij( m+1) between successive
rotational positions m and (m+1), the discretized equation for conservation of mass at the
end and backlash flow areas between time steps t and t + t can be written as

1
t

) ( m)
(em,ij+1)Vij( m+1) (em,ij)Vij( m) + 2 ve( m
,ij Ae,ij = 0,

82

(3.6a)

1
t

( m+1)Vij( m+1) ( m)Vij( m) + v ( m) A( m) = 0.


b,ij
b ,ij b ,ij
b,ij

(3.6b)

(m)
m)
, (e,ij
and Vij( m) are the backlash area discharge
In the above set of equations, b,ij

density, the end area discharge density and the volume of the fluid inside control volume
( m+1)
m +1)
H ij( m) , which is at an incremental rotational position m, b,ij
and (e,ij
and Vij( m) are

the physical quantities for the same control volume H ij( m+1) , but which is now at an
)
( m)
incremental rotational position (m+1), is the ambient air density, and ve( m
,ij and vb,ij

represent the end flow and backlash flow velocities at rotational position m. Simplifying
this further, with t = i i , where i is the incremental rotation angle between
positions m and (m+1) of gear i and i is the rotational velocity of gear i, expressions for
the backlash and end flow velocities can be obtained as

)
vb( m
,ij

i b( m,ij)Vij( m) b( m,ij+1)Vij( m+1)


,
=
( m)
Ab,ij i

(3.7a)

)
ve( m
,ij

i (em,ij)Vij( m) (em,ij+1)Vij( m+1)


.
=
( m)
2 Ae,ij i

(3.7b)

Denoting the energy per unit mass inside the control volume at the end and
backlash flow areas during the rotational increment m by (em,ij) and b( m,ij) , and energy per
unit mass of the same control volume at location (m+1) by (em,ij+1) and b( m,ij+1) , and
83

denoting the energy per unit mass at ambient conditions as , the discretized form of
the second law of thermodynamics [3.7] can be applied to the control volumes between
time instants t and t + t (i.e. between positions m and (m+1)) at the end and backlash
flow areas, respectively, as

( m +1) ( m+1) ( m+1)


(em,ij) (em,ij)Vij( m )
Qh Ws e,ij e,ij Vij
) ( m)

=
+ 2 ve( m
,ij Ae,ij = 0 ,
t
t
t

(3.8a)

( m +1) ( m+1) ( m+1)


b( m,ij) b( m,ij)Vij( m )
Qh Ws b,ij b,ij Vij
) ( m)

=
+ vb( m
,ij Ab ,ij = 0 .
t
t
t

(3.8b)

In the above equation, the energy per unit mass is given by the sum of the internal and
kinetic energies. The thermodynamic process is assumed to be adiabatic ( Qh t = 0 ),
and the shaft work is deemed to be negligible ( Ws t = 0 ). In addition, potential energy
and surface tension effects as well as electrical and magnetic effects are neglected.
Given the fact that there exists no kinetic energy inside the control volume at a
given time instant because of zero velocity inside the control volume, the energy per unit
mass inside the control volume at the end and backlash flow areas during the rotational
)
( m)
increment m is written in terms of the internal energy ue( m
,ij and ub,ij , with the internal
+1)
+1)
energy of the same control volume at position (m+1) given by ue( m
and ub( m
for the
,ij
,ij

respective discharge areas. Denoting the energy per unit mass at ambient conditions as
the sum of the internal energy at ambient conditions, u , and the kinetic energy of the
84

fluid escaping through the end and backlash flow areas, Eqs. (3.8a) and (3.8b) can be
written based on these assumptions in the following form:

+1) ( m +1)
) ( m)
e( m,ij+1)ue( m
(em,ij)ue( m
,ij Vij
,ij Vij

t
+1) ( m +1)
b( m,ij+1)ub( m
b( m,ij)ub( m,ij)Vij( m )
,ij Vij

) 2
(ve( m

,ij )
) ( m)
+ 2
+ u ve( m
,ij Ae,ij = 0 ,
2

(3.9a)

(v ( m ) ) 2

b ,ij
) ( m)
+
+ u vb( m
,ij Ab ,ij = 0.
2

(3.9b)

)
( m)
( m +1)
+1)
)
( m)
( m +1)
Noting that ue( m
= cvTe(,m
, ub( m
= cvTb(,mij +1) and
,ij = cvTe,ij , ue,ij
ij
,ij = cvTb ,ij , ub,ij
)
( m +1)
u = cvT , where cv is the specific heat at constant volume, and Te(,m
, Tb(,mij ) and
ij , Te,ij

Tb(,mij +1) refer to the temperatures corresponding to successive rotational increments, m


and m + 1 , along the end and backlash flow areas, Eqs. (3.9a) and (3.9b) can be written as

+1) ( m +1)
) ( m)
cv [(em,ij+1)Te(,m
(em,ij)Te(,m
ij Vij
ij Vij ]

t
cv [b( m,ij+1)Tb(,mij +1)Vij( m +1) b( m,ij)Tb(,mij )Vij( m ) ]
t

) 2
(ve( m

,ij )
+ 2
+ cvT ve( ,mij) Ae(,mij) = 0 , (3.10a)
2

(v ( m ) ) 2

b ,ij
) ( m)
+
+ cvT vb( m
,ij Ab,ij = 0 . (3.10b)
2

Further, assuming that the thermodynamic process is isentropic, the following relation
can be written for the temperature, pressure and density between two successive
rotational increments [3.7]:

85

+1)
Pe(,m
ij
)
Pe(,m
ij

+1) 1 ( m +1)
Te(,m
e,ij
ij
= ( m)
= ( m) .
Te,ij
e,ij

(3.11)

Here = c p cv is the compressibility factor. With this, Eq. (3.10a) and (3.10b) can be
written in terms of the end and backlash flow velocities and the end and backlash
densities, respectively.
Substituting for the flow velocity along backlash and end flow areas into Eq.
(3.10a) and (3.10b), two non-linear algebraic equations are obtained, one for each
discharge density, which can be solved using the Gauss-Newton method with a suitably
defined step size. Once the density at the backlash and end flow areas are determined
through successive rotational increments m [1, M ] , the pressure at the backlash and end
flow areas can be calculated by two different ways.

The first method uses the

compressible version of the Bernoulli equation [3.7] as

Pb(,mij +1)

( m +1)
1 1 ( m +1) ( m ) 2
b ,ij
= ( m ) Pb(,mij ) +
b,ij (vb,ij ) (vb( m,ij+1) )2 ,

b,ij
2

(3.12a)

+1)
Pe(,m
ij

e( m,ij+1)
) 1 1 ( m +1) ( m ) 2
( m +1) 2
= ( m) Pe(,m
ij +
e,ij (ve,ij ) (ve,ij ) .
e,ij
2

(3.12b)

The second way is to use Eq. (3.11) directly to write

86

Pb(,mij +1)

+1)
Pe(,m
ij

( m+1)
( m ) b ,ij
= Pb,ij
,
b( m,ij1 )

( m+1)
( m ) e,ij
Pe,ij ( m)
e,ij

(3.13a)

(3.13b)

At low Mach numbers, the compressible Bernoulli formulation for pressure


calculations is compatible with the isentropic relationship for pressure computation. The
simulations carried out in Section 3.4 as well as the simulations used in Chapter 4 for the
validation of the windage pocketing power loss model will make use of the isentropic
relationship between density and pressure by computing the pressure inside the control
volume for the backlash and end flow areas using Eq. (3.13a) and (3.13b), respectively.
The rest of the windage pocketing power loss formulation follows the same
procedure as laid out for incompressible fluid flow pocketing formulation of Chapter 2.
The conservation of momentum [3.7] gives the forces acting on the air or air-oil mixture
exiting through the backlash and end flow areas, respectively, as the product of the mass
flow rate times the velocity of fluid ejected through the corresponding discharge areas,
i.e.,

Fb(,mij +1)

+1) ( m+1)
Pb(,mij +1) Ab( m
v
,ij
b,ij
=
RTb(,mij )

P (m)
b,ij
Pb(,mij +1)

87

1
)

(3.14a)

+1)
Fe(,m
=
ij

(
+1)
ve( m
(m)
,ij Pe,ij
+1)
)
Pe(,m
RTe(,m
ij
ij

+1) ( m+1)
Pe(,m
Ae,ij
ij

1
)

(3.14b)

where R = 8.314 J/kg/mol is the universal gas constant and P is the ambient pressure.
Finally, the power loss due to pocketing of control volume H ij( m) at a given incremental
rotational position m is equal to the product of the discharge velocity and the pocketing
force, with the factor of two accounting for both the end flow areas:

( m)
( m) ( m)
(m ) ( m)
Pwp
,ij = vb,ij Fb,ij + 2ve,ij Fe,ij .

(3.15)

This represents the power loss for the j-th control volume of gear i at the m-th rotational
position. The same procedure is applied to all J1( m ) control volumes of gear 1 and J 2( m )
control volumes of gear 2 shown in Figure 2.3 to determine the total instantaneous power
loss of a gear pair due to pocketing, which is averaged over the entire gear mesh cycle to
find the average windage pocketing power loss, very much similar to the way it is defined
in Chapter 2 for the incompressible fluid flow analysis as

Pwp

1
=
M

J 2( m )
J1( m )

(
)
( m)
m

P
+
P
wp,1 j wp,2 j .
m =1 j =1
j =1

88

(3.16)

In the formulation presented above, it is assumed that opposing control volumes


linked to each other through the backlash discharge areas are independent entities, with a
pressure differential existing only between the inside of the control volume and the
ambient conditions. In reality, as explained in the introductory portion of the chapter,
there is the possibility of a pressure differential existing between opposing control
volumes themselves, with the backlash discharge area acting as the bridge in
communicating this pressure differential between the opposing control volumes. The
involved task of considering the pressure differential between opposing control volumes
as dependent entities falls beyond the scope of this dissertation.

3.3 Windage Drag Power Losses


Drag loss formulations of the oil churning model proposed in Chapter 2 are
applicable to the windage power loss model as well, though with some minor changes to
accommodate for the presence of air or air-oil mixture surrounding the gears. The drag
power loss is broken down into power loss taking place on the periphery (circumference)
of the gears and power loss taking place along the faces (sides) of the gears. The
formulations for drag loss on the periphery and faces of the gears are applied to both
gears in the pair, and the total losses are summed up to find the total drag power loss of
the gear pair, in the manner as laid out in Chapter 2.
In applying the periphery drag loss formulation, laminar flow between concentric
rotating disks is considered to be the flow methodology. This represents the scenario of
the outer disk placed at infinity, aping the presence of the flanges/casing, and the inner
89

disc placed at the center mimicking the presence of the gear. Such an approach was used
to come up with an expression for the shear stress built up due to resistance to fluid flow
along the periphery in Chapter 2.

This resistance to fluid flow takes place in the

tangential direction along the periphery of the gears. The drag loss on the periphery is
then given by the product of the built up shear force and the velocity of the fluid along
the periphery of the gear in the tangential direction. Referring to Chapter 2 for equations
concerning peripheral drag due to oil churning, as well as setting the immersion depth
parameter hi = hi roi = 2 where hi is the static depth of the fluid around the gear i and

roi is the outside (major) radius of gear i (i.e. gear is fully submerged in air or air-oil
mixture), the windage drag power loss of gear i along its periphery is derived as

Pwdpi = 4 bi roi2 i2 .

(3.17)

Here, is the dynamic viscosity of air (or a homogeneous mixture of air and oil) and bi
is the face width of gear i.
Similarly, the power losses due to air drag along the faces of gear i can be derived
from the same oil churning formulations of Chapter 2. In formulating the drag loss along
the faces of the gears, a boundary layer approach was used in Chapter 2, with the laminar
and turbulent regimes handled separately, based on the Reynolds number for fluid flow
across the faces. Also, in Chapter 2, a linear velocity profile was considered for fluid
flow across the faces of the gears to represent laminar conditions, and the boundary layer
thickness and skin friction coefficient were calculated through the application of the von
90

Karman momentum integral equation [3.7]. Finally, the power loss due to drag on the
faces of gear i was written as the product of the shear stress built up across the faces of
the gear due to boundary layer formation and the velocity of the fluid flowing across the
faces of the gear. With h = 2 , windage drag power loss of gear i along its faces in the
laminar regime is hence given as

( L)
2.5 4
Pwdfi
= 0.41 0.5
k i roi ,

(3.18)

where superscript L denotes the laminar regime, and k and are the kinematic
viscosity and density of air (or air-oil mixture). Likewise, the oil churning power loss
model of Chapter 2 used the Prandtl-one-seventh power law for the velocity profile in
turbulent regime, from which the boundary layer thickness and skin friction coefficient
were computed again, using von Karmans equation [3.7]. The turbulent flow face drag
formula is adopted here for computing the windage power loss across both the faces of
the gear, with flow being turbulent and by setting h = 2 , as

(T )
Pwdfi
= 0.0250.14
i2.86 roi4.72 .
k

(3.19)

3.4 Example Windage Power Loss Analyses and Parametric Studies

In order to investigate the influence of various operating parameters and also to


quantify the contributions of individual components on the total windage power loss, two
91

unity-ratio example spur gear pairs, whose gear geometry parameters are listed in Table
2.1 (for the sake of consistency, the same family of spur gear pairs used in the oil
churning analysis of Chapter 2 is also used for the example analysis presented in this
chapter), are considered for this purpose. It is to be noted that the backlash flow area is
given by the product of the face width of the gears and the backlash length at all points
over one mesh cycle, keeping the variations in backlash flow area consistently the same
throughout the mesh cycle. This is considered for the sake of simplicity of analysis. The
central theme of this section is then to (i) demonstrate the effective contributions of drag
and pocketing power loss components to the total windage power loss Pw and (ii)
investigate the influence of various gear pair parameters on Pw .
The first example gear pair is a 40-teeth design with a module of m% = 2.32 mm
and three different face width values of b = 14.7, 19.5 and 26.7 mm , respectively. The
second example spur gear pair is a 23-teeth design, with m% = 3.95 mm and b = 19.5 mm .
Both gear pairs are operated at a center distance of 91.5 mm, under zero transmitted
torque and under jet lubrication conditions. In these example analyses, the rotational
speed = 1 =

60
2 1

is varied within a range of up to 10,000 rpm (corresponding to a

maximum gear pitch line velocity of about 48 m/s).


Figure 3.1(a) shows the influence of temperature on Pw for the first example gear
pair (40-teeth) with m% = 2.32 mm and b = 19.5 mm. The operating gearbox temperature
ranges from 30o C to 100o C .

For computing the windage drag power losses, the

properties

respective

of

air

at

those

92

temperature

values

were

considered.

1.6
(a)

1.2

(b)

T = 30 o C

b = 26.7 mm

T = 50 o C

b = 19.5 mm

T = 70 C

b = 14.7 mm

Pw

T = 100 C
o

0.8

[kW]

0.4

0.0
0

2000

4000

6000

8000

10000 0

2000

4000

6000

8000

10000

1.6
(c)

m% = 3.95 mm
1.2

m% = 2.32 mm

Pw
[kW]

0.8

0.4

0.0
0

2000

4000

6000

8000

10000

[rpm]

Figure 3.1 Effect of (a) temperature T , (b) face width b and (c) gear module m% on Pw .

93

Windage pocketing power loss was computed by evaluating the density at the backlash
and end flow areas at a particular operating temperature value. As seen in Figure 3.1(a),
Pw increases with reduction in temperature, regardless of the operating speed, due to a

significantly non-linear influence of the temperature on Pwp and a weakly non-linear


effect of temperature on the drag power losses Pwd . As a sample quantification of the
effect of temperature on the windage power losses, Pw at 30o C and 7,000 rpm is
predicted to be about 0.36 kW higher that Pw value at the same speed and at 100o C .
This is a direct result of the influence of temperature on both viscosity and density of the
medium surrounding the gears as well as inside the meshing zone. As the operating
temperature increases, the air/air-oil mixture inside the gearbox, being compressible in
nature, sees an increase in the operating kinematic viscosity value and a decrease in
operating density. This combined effect in turn causes a decrease in the windage power
loss value, as evinced by the curves in Figure 3.1(a), with density having a stronger
influence in reducing the windage power loss than kinematic viscosity.
Figure 3.1(b) shows the influence of face width b on Pw for a gear pair having
m% = 2.32 mm operating at 80o C . The gear pair having b = 26.7 mm is predicted to

experience about 1.5 times the total windage power loss as compared to a gear pair with
b = 19.5 mm, and about 2.5 times the total windage power loss as experienced by the

gear pair at b = 14.7 mm, all at the same operating speed and temperature conditions.
For instance, Pw for the gear pair having wide face width of b = 26.7 mm at 80o C and
5,000 rpm is 0.22 kW, as compared to Pw of 0.14 kW for the gear pair with medium face
94

width b = 19.5 mm, and Pw of 0.09 kW for the gear pair with narrow face width b = 14.7
mm, all at the same operating temperature and speed values. An increase in face width
opens up more of the control volume for ejection of the fluid and as a result, a greater
velocity profile can be predicted, leading to higher windage pocketing power losses and
hence, total windage power loss.
In order to throw more light on the influence of face width on windage power
loss, windage pocketing and drag power loss for the gear pair with m% = 2.32 mm and
varying face width values are shown in Figures 3.2(a) and 3.2(b), respectively, at the
same operating conditions. Corresponding to the formulation laid out earlier in Chapter
2, the influence of b on windage pocketing power loss is significant, very much more so
than it is on the windage drag power loss. It is also seen in Figure 3.2 that the pocketing
loss of this particular gear pair constitutes a very large portion of Pw (in this example,
more than 98%) with drag losses more or less staying as a tertiary influence under these
operating conditions. Hence, the variation of the total windage power loss with face
width is primarily due to the windage pocketing power loss, which is strongly affected by
changes in face width.
Figure 3.3 shows the individual components of windage pocketing power loss, in
terms of power loss through the backlash and end flow areas, for varying face widths and
at an operating temperature of 80o C over the same speed range. As seen in this figure,
the value of the face width has a strong influence on both the end and backlash
components of windage pocketing power loss, with an almost two and a half fold
increase taking place between windage pocketing power loss over narrow and wide face
95

1.0
(a)

b = 26.7 mm

0.8

b = 19.5 mm
Pwp

b = 14.7 mm

0.6

[kW]

0.4

0.2

0.0
0

2000

4000

6000

8000

10000

6000

8000

10000

0.008
(b)

0.006

Pwd

0.004

[kW]

0.002

0.000
0

2000

4000

[rpm]

Figure 3.2 Components of Pw for the gear pairs having m% = 2.32 mm and varying b at

80o C ; (a) windage pocketing power loss Pwp (b) windage drag power
losses Pwd .
96

1.0
(a) b = 14.7 mm

(b) b = 19.5 mm

Pwp

0.8

Pwp, e
Pwp,b
Pwp

0.6

[kW]

0.4

0.2

0.0
0

2000

4000

6000

8000

10000 0

2000

4000

6000

8000

10000

1.0
(c) b = 26.7 mm

0.8

Pwp

0.6

[kW]

0.4

0.2

0.0
0

2000

4000

6000

8000

10000

[rpm]

Figure 3.3 Windage power loss Pwp and its components, Pwp,e and Pwp,b , for gear pairs

having m% = 2.32 mm at 80o C ; (a) b = 14.7 mm (b) b = 19.5 mm (c)


b = 26.7 mm.
97

width values.

Further portrayed in the figures is the constitution of the windage

pocketing power loss in terms of its individual components. At narrow face width values
( b = 14.7 mm), the backlash and end power loss components of the windage pocketing
power loss contribute more or less equally to the total windage pocketing power loss, as
seen in Figure 3.3(a). As the face width is increased, the end pocketing power loss is the
more dominant component of the total windage pocketing power loss, with pocketing
power loss at the ends constituting over 65 % of the total pocketing power loss over the
entire operating speed range at wide face width values, seen in Figure 3.3(c).
The final comparison in Figure 3.1(c) portrays the influence of m% on Pw for
b = 19.5 mm at 80o C .

The 23-teeth coarse pitch gear pair, with m% = 3.95 mm, is

predicted to have more windage power loss than the 40-teeth gear pair having m% = 2.32
mm. As an example, at 5,000 rpm, Pw for m% = 3.95 mm is about 0.15 kW greater than
that of the 40-tooth gear pair having m% = 2.32 mm. As Pwdf and Pwdp components are
not expected to be influenced significantly by tooth size according to the formulations
presented, this difference in Pw values with module can be attributed primarily to Pwp .
This is due to the fact that volume change for the coarser pitch gear pair is greater than
for the fine pitch 40-teeth gear pair. Since volume change depends on the size of the
mating teeth falling into place in the tooth space cavity as the gears rotate, a larger tooth
can cause greater volume displacement as tracked by the 23-teeth coarse pitch gear pair
and hence, a greater pocketing power loss is attributed to this fact, both under
compressible and incompressible flow conditions.

98

The variation of windage power loss with operating speed is non-linear in nature,
as seen in Figures 3.1 - 3.3. For instance in Figure 3.2, for b = 19.5 mm, Pw 12.1
while the power loss-speed relationships for individual components are Pwdf 2.5 (or

Pwdf 2.86

for turbulence), Pwdp 2 and Pwp 1.8 .

The trend of windage

pocketing power loss variation with speed is mixed and non-linear in nature. At lower
operating speed values, the windage pocketing power loss varies with speed in a
quadratic fashion, similar to the one depicted in incompressible flow conditions as seen in
Chapter 2. At higher operating speeds, the windage pocketing power loss shows an
almost linear behavior, with the steeper rise in the velocity of fluid being ejected
contributing as a direct effect towards this trend.
Figures 3.4, 3.5 and 3.6 are included here in order to throw more light on the
physics of the phenomenon of windage pocketing power losses. In Figure 3.4(a) and
3.4(b), the variation of the end flow area of a single control volume with gear rotation is
shown for the gears having m% = 2.32 mm and m% = 3.95 mm, respectively. The range of
end area values shown for the fine pitch and coarse pitch gear pair has to be understood
from the point of view of the total cavity area available for the pocketing purposes. As an
average, the total cavity area between adjacent teeth for the 40-teeth gear is about
11.5 mm 2 (roughly, the product of the working tooth height times the average space
between adjacent teeth, which is given by the difference between the circular pitch and
the tooth thickness at pitch point), and for the 23-teeth gear pair, the total cavity area
afforded for pocketing purposes is approximately 50.6 mm 2 .
99

30

(a) m% = 2.32 mm

25

( m)
H11

20

H12

( m)

m)
Ae(,1
j

( m)
H10

15
[mm2 ]

10

5
0

30

(b) m% = 3.95 mm

25

20
m)
Ae(,1
j

15

[mm2 ]

10

0
0

Rotational Position (m)

Figure 3.4 Variation of the end flow area Ae(,1m)j with rotational position for the gear pairs

having (a) m% = 2.32 mm , and (b) m% = 3.95 mm.


100

From Figure 3.4(a), the variation of the end flow area corresponding to the pocket
(m)
H11
shows that the end area decreases as the gears enter into mesh with each other,

reaching a minimum position at the mating pitch point and then expanding beyond that
(m)
point. At the same time as the control volume H11
undergoes compression/expansion,
(m)
depending on the contact ratio, there will be other volumes, for example H12
and
(m)
H10
in Figure 3.4(a), which will be undergoing compression and expansion,

simultaneously, with consecutive contact points separated by one base pitch distance. In
this way, over one base pitch rotation of the gears, all the pockets undergoing
compression and expansion have to be considered for computing the average pocketing
power loss, as explained in the methodology for power loss computation. Similar such
variations of end flow area is encountered in case of the other example gear set, with
compression/expansion taking place over different intervals for the 23-teeth gear pair due
to differences in module, contact ratio and number of teeth, and hence, cavity areas. As
explained earlier, the backlash discharge area is given by the product of the backlash and
the face width values for the respective sample gear pairs, and hence, variations in
backlash area is neglected for the sake of simplicity.
Figure 3.5 shows the variation of density, pressure and velocity along the end and
(m)
backlash flow areas for control volume H11
of the 40-teeth gear pair ( m% = 2.32 mm,

b = 19.5 mm) at = 1,000, 5,000 and 10,000 rpm. Figures 3.5(a1), 3.5(b1) and 3.5(c1)

show the variation of density, pressure and velocity along the end flow area with respect
(m)
while Figures 3.5(a2), 3.5(b2) and 3.5(c2) show the same
to rotational position for H11

101

(a1)

1,000 rpm

5,000 rpm

( m)
e,11

( m)
b,11

10,000 rpm

4
[kg/m3 ]

(a2)

[kg/m3 ]

1
1.0

1.0

(b1)

0.7

(b2)

0.7

)
pe(m
,11

( m)
pb,11

[GPa]

[GPa]
0.4

0.4

0.1

0.1
100

100

(c1)

(c2)

50

50

)
ve(m
,11

(m)

vb,11

[m/s]

[m/s]

-50

-50

-100

-100

Rotational Position (m)

Rotational Position (m)

(m)
Figure 3.5 Variation of (a) density, (b) pressure and (c) velocity of control volume 11
of the gear pair having m% = 2.32 mm with rotational position for (a1, b1, c1)

the end area and (a2, b2, c2) the backlash area ( b = 19.5 mm , 80o C ).

102

(a1)

(a2)

1,000 rpm

5,000 rpm

)
e(m
,11

)
b(m
,11

10,000 rpm

4
[kg/m3 ]

[kg/m3 ]

1
1.0

1
1.0

(b1)

0.7

(b2)

0.7

( m)
pe,11

)
pb( m
,11

[GPa]
0.4

[GPa]
0.4

0.1
100

0.1
100

(c1)

50

(c2)

50

)
ve( m
,11

)
vb( m
,11

[m/s]

[m/s]

-50

-50

-100

-100
0

Rotational Position (m)

Rotational Position (m)

(m)
Figure 3.6 Variation of (a) density, (b) pressure and (c) velocity of control volume 11
of the gear pair having m% = 3.95 mm with rotational position for (a1, b1, c1)

the end area and (a2, b2, c2) the backlash area ( b = 19.5 mm , 80o C ).

103

along the backlash flow area. As seen from the figures, the variation of the governing
parameters over one mesh cycle follows the compression/expansion process for the
respective pocket and flow areas over the same mesh cycle. The velocity of the fluid
ejected through the end and backlash flow areas, as seen in Figures 3.5(c1) and 3.5(c2),
respectively, shows negative values for velocity just beyond the point in the mesh cycle
where expansion has started taking place, whereby fluid is sucked into the pocket due to
the loss of pressure differential created by the expansion of the control volume. Variation
of density and pressure with rotational position for the backlash flow area shows a
slightly higher value than those for the end flow area. This is due to the fact that the
backlash flow area has been assumed, for the sake of computational simplicity, to be the
product of the face width and the backlash length, while the end flow area has been
allowed to vary, as seen in Figure 3.4(a) for the 40-teeth gear pair. Over the entire mesh
cycle, the average end flow area is larger in value than the backlash flow area, resulting
in lower density values for the end flow area versus the backlash area. Correspondingly,
the pressure changes with rotational position, for the end and backlash flow areas, in
Figures 3.5(b1) and 3.5(b2), respectively, show very similar trends and within the same
operating range, for both the flow areas.
Also noted in Figure 3.5 is the variation of these parameters with speed,
validating the physical notion that an increase in speed corresponds to an increase in the
density and pressure and hence, an increase in the velocity of the fluid being squeezed
out. The variation of velocity with respective flow areas do not show a similar trend with
respect to one another because of the way the velocity is computed. The difference
between product of volume and density changes between rotational positions are
104

considered in the computation of fluid velocity for the end and backlash flow areas, as
given by Eq. (3.7), and hence, steeper changes in such differential products between
mesh positions will produce steeper velocity gradations, as seen in Figures 3.5(c1) and
3.5(c2).
As reported in Figure 3.5, similar such trends are noted in Figure 3.6, where the
same parameters are presented for the 23-teeth coarse pitch gear pair ( m% = 3.95 mm,
b = 19.5 mm). Evidently, the coarser pitch gear pair has a higher velocity, density and

pressure differential than the fine pitch 40-teeth gear pair, due to larger changes taking
place with respect to volume variation within mesh position over each incremental step,
hence leading to greater windage pocketing power losses.
It is also worthwhile to note that each analysis performed and outlined in this
study took less than a second of CPU time, corroborating the fact that the proposed
windage power loss model eschews the computational intensity of CFD type
formulations, making it viable as an effective and practical design tool.

3.5 Concluding Remarks

A physics-based fluid mechanics power loss model was proposed to predict power
losses for gear pairs operating under windage conditions. The framework of the model
included individual formulations for windage losses on the periphery and faces of the
gears as well as a compressible fluid model for power loss due to pocketing taking place
in the meshing zone.

The windage conditions simulate jet-lubricated operating


105

conditions or low oil-level dip-lubricated conditions. As the first stab, the windage
power loss model was applied to two example unity-ratio gear sets with varying gear
geometry parameters, to quantify the contributions of each of the components of the total
windage power loss. In both example cases, the windage pocketing losses were shown to
dominate the total gear pair windage losses, as was the case in incompressible flow
computations performed in Chapter 2. Also, the influence of operating conditions, gear
geometry parameters and lubricant properties on windage power loss was quantified for
the gear pairs in consideration.

106

References for Chapter 3

[3.1]

Petry-Johnson, T. T., Kahraman, A., Anderson, N.E., and Chase, D. R., 2008, An
Experimental Investigation of Spur Gear Efficiency, ASME Journal of
Mechanical Design, 130 (6), 062601, 10 pages.

[3.2]

Dawson, P. H., 1984, Windage Loss in Larger High-Speed Gears, Proceedings


of the Institution of Mechanical Engineers, Part A: Power and Process

Engineering, 198(1), 5159.


[3.3]

Diab, Y., Ville, F., and Velex, P., 2006, Investigations on Power Losses in High
Speed Gears, Journal of Engineering Tribology, 220, 191298.

[3.4]

Eastwick, C. N., and Johnson, G., 2008, Gear Windage: A Review, ASME
Journal of Mechanical Design, 130, 034001 (6 pages).

[3.5]

Wild, P. M., Dijlali, N., and Vickers, G. W., 1996, Experimental and
Computational Assessment of Windage Losses in Rotating Machinery, ASME
Trans. J. Fluids Eng., 118, 116122.

[3.6]

Al-Shibl, K., Simmons, K., and Eastwick, C. N., 2007 , Modeling Gear Windage
Power Loss From an Enclosed Spur Gears, Proceedings of the Institution of
Mechanical Engineers, Part A, 221(3), 331341.

[3.7]

Streeter, L. V., and Wylie, E. B., 1985, Fluid Mechanics, Eighth edition,
McGraw-Hill, New York.
107

CHAPTER 4

VALIDATION OF OIL CHURNING AND WINDAGE


POWER LOSS MODELS

4.1 Introduction
In the previous two chapters, two novel models, one for predicting oil churning
power losses, and the other for predicting windage power losses of spur gear pairs, were
proposed. Both the models employed fundamental principles of fluid mechanics and a
host of critical assumptions, so that the formulations were reduced to closed-form
expressions, making them practical for application towards design purposes. In cases
where a numerical computational methodology has to be used, care was taken to rein in
the computational effort, as is the case for the compressible fluid pocketing power loss
component of the windage power loss model. While these two models are desirable in
the sense that they can be used conveniently and efficiently to predict power losses
associated with gear pairs, they would not be viable tools for actual product development
unless they are validated through real-world experiments conducted on gear pairs. This
chapter takes on the onerous task of such a validation of both the power loss models.

108

It has to be noted upfront that the idea of such a validation process is not to
provide for accurate point-to-point comparison with experiments; such a comparison is
not possible when experimental uncertainties and modeling assumptions are taken into
account. The broader objective is to provide for a comparison between predictions and
experiments, wherein the predictions follow the experimental trend, within limits of
experimental scatter, over wide ranges of operating parameters. Provided the models can
produce such predictions, they will be deemed validated as practical and computationally
efficient tools, which can provide a reasonable estimate of load-independent power losses
occurring in gear pairs.
The most basic gear system, a spur gear pair, will be used in the experimental
comparisons in this chapter. Another system-level validation will be provided in Chapter
5, where the proposed spin loss models will be applied to a six-speed automotive
transmission and its predictions will be compared to transmission power loss
measurements.
As a part of this study on load-independent losses, a companion experimental
study was performed by Moorhead [4.1] to measure oil churning power losses of unityratio spur gear pairs having different design parameters.

These experiments were

performed at different static oil levels and operating oil temperature, as well as over a
wide range of operating speed. The oil churning power loss model of Chapter 2 will be
used to simulate the experiments of Moorhead [4.1] for verifying the trends exhibited by
the predictions as well as for quantifying the accuracy of the predicted power loss values

109

through a direct comparison to experimental results. Section 4.2 will briefly describe the
experimental set-up of Moorhead [4.1] and provide comparisons between his experiments
and model predictions for validating the oil churning power loss model.
For validation of the windage power loss model of Chapter 4, the experimental
database of Petry-Johnson et al [4.2] will be used. This body of experimental data was
generated by performing load-independent tests on jet-lubricated spur gear pairs, with no
accumulation of the oil within the gearbox, so as to simulate windage conditions at high
operating speeds. These experiments will be simulated in Section 4.3 to determine the
accuracy of the windage power loss model predictions. In addition, a direct comparison
of the windage model formulations to the empirical formula of Dawson [4.3] will also be
carried out as a part of this validation effort.

4.2 Validation of the Gear Pair Oil Churning Power Loss Model
4.2.1 Test Machine and Oil Churning Test Procedure
A gear efficiency test machine used previously by Petry-Johnson et al. [4.2] to
study power losses of jet-lubricated gearboxes, as shown in Figure 4.1(a), was used by
Moorhead [4.1], with some modifications, to perform gear pair oil churning experiments.
Only the most relevant details of the test set-up are provided here. The test machine
consists of two identical gearboxes, each containing one gear pair and four identical
cylindrical roller bearings. In this power circulatory arrangement, one gear from each
gearbox is connected to the corresponding gear of the other gearbox, as shown
110

test gearboxes

torque meter

belt drive

1
high-speed spindle

AC motor

Split Load
Coupling

Figure 4.1 (a) A view and (b) the layout of the gear efficiency test machine [4.1, 4.2].

111

schematically in Fig. 4.1(b) [4.1]. The input shaft of the reaction gearbox is connected to
a high-speed spindle through a flexible coupling and a precision non-contact type torquemeter. The high-speed spindle is driven by a belt with a 3:1 ratio speed increase from a
variable speed AC motor.
Figure 4.2 shows a picture of one of the gearboxes that is configured to perform
dip-lubricated efficiency tests [4.1].

The other gearbox has the exactly the same

arrangement. Valves of the oil return were closed in order to maintain an oil bath within
each gearbox. Equal static oil levels were ensured in both the gearboxes via a flow path
between them. Sheet metal baffle plates were added inside the gearboxes in front of the
oil drainage ports to minimize oil circulation in the drainage lines. Large diameter
graduated site glasses were used to monitor the static oil level within each gearbox. A
ball valve placed between each gearbox and its site glass was closed during the tests to
ensure that the site glasses did not act as fluid capacitors.
The power circulation loop can be activated by applying a torque at the split
coupling to load the gear meshes. Here, only a very little amount of such a torque was
applied so that any tooth separations or rattle can be prevented. In addition, all the test
gears were super-finished through a chemical polishing process to minimize the surface
roughness amplitudes.

This way, the load-dependent (mechanical) losses in these

experiments can be deemed negligible and the total torque provided to the loop, TT , as
measured by the torque-meter, represents the total spin losses from both gearboxes.
Since both gearboxes are identical, the spin power loss of an individual gearbox is

112

Figure 4.2 One of the test gear boxes shown in dip-lubrication arrangement [4.1].

113

determined as PsT = 12 TT , where is the rotational speed of the unity-ratio test spur
gears.
Torque TT was measured with a non-contact digital torque-meter. Moorhead
[4.1] reported that the torque-meter has a full-scale range of 50 N-m, and overall
accuracy of 0.03% of full scale for the installed system. It was also reported that the
rotational speed = 1 = 2 = 60 2 (in rpm) was typically maintained within 0.20.3% of the set (intended) speed, and exhibited a negligible ( 2 rpm) fluctuation about
this value, which was averaged out during data collection [4.1].
4.2.2 Gear Specimens and Parameters Studied
Two basic unity-ratio spur gear designs were implemented in Moorheads study
[4.1]. The design parameters of these test gears were provided earlier in Table 2.1.
Figure 4.3 shows a picture of a pair of these test gears [4.1]. Moorheads study [4.1]
provided data for investigating the influence of the following parameters:
(i)

Oil Viscosity Oil churning tests were performed by using a typical manual
transmission fluid. The lubricant temperature was not controlled, and was allowed
to vary while being recorded continuously [4.1].

Each test was started with

lubricant temperature at 30C, and spin losses were measured as oil temperature
rose gradually to its steady-state value. Then, another test was started at the same
test conditions, but now with the oil heated to an elevated temperature (say 100C)
and the test continued until the gearbox was cooled down to the same steady-state

114

Figure 4.3 Two examples of test gears (a) 23-tooth gear with m% = 3.95 mm and
b = 19.5 mm, and (b) 40-tooth gear with m% = 2.32 mm and b = 19.5 mm
[4.1, 4.2].

115

temperature. Combining these two tests provided continuous spin loss data for each
condition within a wide range of oil temperature from 30C ( k = 131.2 cSt) and
100C ( k = 13.1 cSt). This allowed for a complete evaluation of the influence of
lubricant temperature and viscosity for every test.
(ii)

Rotational Speed The rotational speed of gears at each test was varied up to 6,000
rpm (corresponding to a pitch-line velocity of 29 m/s).

Either 6-point

( = 1, 000 rpm to 6,000 rpm with an increment of 1,000 rpm) or 3-point


( = 2, 000 , 4,000 and 6,000) tests were performed for each test to generate
sufficient data for quantifying the impact of the rotational speed on spin power
losses.
(iii) Gear Module Two different test gear designs, a 40-teeth gear pair having a module

of m% = 2.32 mm and a 23-teeth gear pair having a module of m% = 3.95 mm were


included in the test matrix to collect data on the influence of gear module.
(iv) Gear Face Width Tests on gears of module m% = 2.32 mm were performed with
three different face width values to investigate the impact of the face width on spin
power loss. The face width values used are b = 14.2, 19.5 and 26.7 mm.

(v)

Immersion Depth Figure 4.4 shows various static oil levels considered for the
actual test gearbox with the side cover removed. According to Figure 4.4(a), these
unity ratio gear pairs with ro1 = ro 2 have the same immersion depth h1 = h2 = h

116

(a)

ro
r

h1 = h2
r

pitch circles
(b)

h = 1.5
r

h = 1.0
r

h = 0.5
r

h = 0.05
r

Figure 4.4 Illustration of oil level parameters [4.2].

117

since gear centers O1 and O2 are on the same horizontal plane.

Under this

configuration,

given

the

dimensionless

immersion

parameter

is

as

h = h1 ro1 = h2 ro 2 . The test matrix included tests conducted at values of h = 0.5,


1.0 and 1.5, as shown schematically in Figure 4.4(b). Here, h = 1.0 indicates an oil
level at the centers of the gears (half immersion), and h = 0.5 and 1.5 represent 25%
and 75% immersion, respectively.

4.2.3 Comparison of Oil Churning Power Loss Model to Experiments


The predictions of the spur gear oil churning power loss model formulated in
Chapter 2 will be compared to the experimental data of Moorhead [4.1] to determine the
accuracy of the oil churning power loss model; all of the comparisons will be done for an
up-in-mesh rotational condition. The comparisons to predictions will be presented as a
function of operating speed, , thus allowing for a better quantitative comparison.
Since the measured total spin power loss, PsT , represents the gearbox spin power losses
that include bearing losses as well, the viscous bearing torque losses for the four
cylindrical element bearings are calculated in mW as [4.4]

Pb,spin

4(107 ) f L k 2 / 35 / 3d m3

6.4(105 ) f L d m3

118

k > 2000
k < 2000

(4.1)

where d m is the mean bearing diameter in millimeters ( d m = 60 mm in the experiments),


f L is a bearing viscous friction constant depending on the type of bearing and lubrication
method ( f L =2.2 - 4 for oil bath type lubrication; in this study, f L = 4 ), is the shaft
speed in rpm, is the shaft speed in rad/s and k is the kinematic viscosity of the
lubricant in mm2/s, and added to the spin power loss values for the gear pair to obtain the
total PsT of the gearbox. One should note that Eq. (4.1) is rather simplistic and does not
include h as a parameter for oil churning effects at the bearing locations. It was used in
this study due to the lack of a more sophisticated published bearing model. The shaft seal
losses were reported to be negligible as compared to bearing power losses [4.2], and
hence, were not included in the total spin power loss predictions.
In Figure 4.5, the comparisons between measured and predicted PsT values of a
gear pair having m% = 2.32 mm and b = 19.5 mm are presented at four different oil
temperature values when h = 1.0 . A wide range of inlet temperatures ranging from 30 to
100C is chosen for this comparison. The total spin power loss decreases with an
increase in oil inlet temperature for both measurements and predictions since the
viscosity reduces with an increase in oil temperature. The decrease in measured power
loss is not as steep as it is for the predictions between temperature ranges of 70C to
100C. It is also seen that the comparison between the measured and predicted PsT
values is reasonably good, with predictions well within 0.1 kW of measurements, with
the exception of the data point corresponding to = 4, 000 rpm at 100C where the
difference is about 0.2 kW. Such deviations at higher speeds and higher operating
119

1.2

(a) k = 131.2 cSt ( 30 o C )

(b) k = 54.3 cSt ( 50 o C )

Experiment
Prediction - Total
Prediction - Bearings

0.8

0.4

0.0
PsT
[kW] 1.2

(d) k = 13.1 cSt (100 o C )

(c) k = 27.8 cSt ( 70 o C )

0.8

0.4

0.0
0

1000

2000

3000

4000 0

1000

2000

3000

4000

[rpm]

Figure 4.5 Comparison of predicted to the measured [4.2] PsT for a gear pair (a) 30o C ,
(b) 50o C , (c) 70o C , and (d) 90o C ; m% = 2.32 mm, b = 19.5 mm and h = 1.0 .

120

temperatures can potentially be attributed to the simplicity of the bearing viscous power
loss model, which does not specifically take into account variations in oil viscosity, as it
considers only the range afforded by the product of oil viscosity and operating speed,
together. It is also observed from Figure 4.5 that the predicted variation of PsT with
also agrees reasonably well with the measured trends.
Figure 4.5 also shows the bearing power loss component of the total predicted
spin power loss at the different temperature levels. The bearing power losses (from all
four rolling element bearings) account for about one-third of the total gearbox losses,
indicating that the remaining two-third comes from the gears. Hence, while bearing
power losses are not negligible, they fail to overshadow the pocketing and drag
components of the gear churning power losses.

Next, the predicted and measured influences of immersion depth parameter h on


PsT are compared in Figure 4.6 for four different oil levels. These comparisons are for a
gear pair with b = 19.5 mm and m% = 2.32 mm at 80C.

The predicted and measured

PsT shown in Figure 4.6 also compare reasonably well with differences being within 0.1
kW, with the exception of the high speed portion of the h = 1.5 curve. At high speeds
and higher oil level immersion, assumptions of static oil level as considered while
modeling the oil churning power loss underestimates the total power loss, shown through
the discrepancy in Figure 4.6(d); swell effects of the oil need to be considered while
modeling the churning losses. Overall, it can be stated that the model captures the

121

1.2

(a) h = 0.05

(b) h = 0.5

Experiment
0.8

Prediction

0.4

0.0
PsT
[kW] 1.2

(d) h = 1.5

(c) h = 1.0

0.8

0.4

0.0
0

1000

2000

3000

4000 0

1000

2000

3000

4000

[rpm]

Figure 4.6 Comparison of predicted vs. measured PsT for a gear pair at (a) h = 0.05 , (b)
h = 0.5 , (c) h = 1.0 , and (d) h = 1.5 ; oil at 80o C
m% = 2.32 mm, and b = 19.5 mm.

122

(up-in-mesh),

influence of h on PsT sufficiently. Figure 4.6 also establishes the significance of h on


PsT clearly, with both measured and predicted PsT values increasing with an increase in
immersion depth. As an example, at an operating speed of = 3, 000 rpm, predicted PsT
is 0.21 kW at h = 0.5 , which is almost doubled to 0.41 kW when h = 1.5 .
Figure 4.7 provides comparisons between measured and predicted PsT values at
three different face width values of b = 14.7 , 19.5 and 26.7 mm, for gears having
m% = 2.32 mm, operated at h = 1.0 and oil temperature of 80C. The measured and
predicted PsT values again agree quite well for all three b values, as maximum difference
is less than 0.1 kW.

This figure shows sizable increase in PsT as the face width

increases. For instance, measured and predicted PsT values are 1.01 kW and 0.98 kW,
respectively, for the gear pair having b = 26.7 mm in Figure 4.7(c) at = 4, 000 rpm,
whereas corresponding values for b = 19.5 mm are 0.72 kW and 0.64 kW in Figure 4.7(b)
that are about 0.3 kW less than those for b = 26.7 mm. Likewise, the measured and
predicted PsT of the pair having b = 14.7 mm at = 4, 000 rpm are 0.53 kW and 0.49
kW representing about half of the losses at b = 26.7 mm. Figure 4.7 also shows clearly
that the model is capable of capturing the influence of b on PsT accurately.
As a final comparison, Figure 4.8 shows the influence of module m% on measured
and predicted PsT . Two separate gear pairs, one with m% = 2.32 mm and the other with
m% = 3.95 mm (both having b = 19.5 mm) are considered for this comparison at h = 1.0

123

1.2

1.2

(b) b = 19.5 mm

(a) b = 14.7 mm

Experiment
0.8

0.8

Prediction

PsT
[kW]
0.4

0.4

0.0
0

1000

2000

3000

0.0
4000 0

1000

2000

3000

4000

1.2

(c) b = 26.7 mm

0.8

PsT
[kW]
0.4

0.0
0

1000

2000

3000

4000

[rpm]

Figure 4.7 Comparison of predicted to the measured [4.1] PsT for a gear pair having
m% = 2.32 mm and face width values (a) b = 14.7 mm, (b) b = 19.5 mm, and
(c) b = 26.7 mm; oil at 80o C and h = 1.0 .

124

1.2
(b) m% = 3.95 mm

(a) m% = 2.32 mm

Experiment
0.8

Prediction

PsT
[kW]

0.4

0.0
0

1000

2000

3000

4000 0

1000

2000

3000

4000

[rpm]

Figure 4.8 Comparison of predicted to the measured [4.2] PsT for a gear pair of face
width b = 19.5 mm and modules (a) m% = 2.32 mm and (b) m% = 3.95 mm; oil
at 80o C and h = 1.0 .

125

(up-in-mesh) and oil temperature of 80C.

The agreement between the model and

experiments are again well within 0.1 kW in both the cases. Gear module comes across
as an important parameter, influencing both the spin and mechanical power loss and
hence, is a key operating parameter in ensuring the reduction of power losses with lower
module.

4.3 Validation of the Gear Pair Windage Power Loss Model


For validation of the windage power loss model, measurements of Petry-Johnson
et al. [4.2] will be compared against the predictions of the windage power loss model
detailed in Chapter 4. Petry-Johnson et al [4.2] used jet-lubrication for most of the
experiments. The test machine set-up and methodology for measuring spin power losses
under jet-lubrication conditions are described in detail in ref. [4.2, 4.5]. The same test
machine as explained in ref. [4.1] was employed in this case as well with the same test
gear pairs whose parameters listed in Table 2.1. The spin power loss measurements were
carried out at rotational speeds of = 2, 000 rpm to = 10, 000 rpm in increments of
2,000 rpm. The 40-teeth and 23-teeth test gears have different face width values, apart
from having different modules. The operating temperature was maintained at 110C.
Figure 4.9 provides for direct comparisons between predictions of the windage
power loss model of Chapter 3, and measured data from Ref. [4.2]. First, predicted and
measured windage power loss values, Pw , for the 40-teeth gear pair having m% = 2.32
mm and b = 26.7 mm are compared in Figure 4.9(a). Here, the gearbox bearing losses
are accounted for by using the same viscous bearing power loss formula [4.4],
126

1.6
Experiment

Prediction
1.2

Pw

0.8

[kW]

0.4

0.0
0

2000

4000
6000
[rpm]

8000

10000

2000

4000

8000

10000

1.6

1.2

Pw

0.8

[kW]

0.4

0.0
6000

[rpm]

Figure 4.9 Comparison of predicted and measured [4.2] Pw for (a) the 40-tooth gear
pair having m% = 2.32 mm and b = 26.7 mm and (b) 23-tooth gear pair
having m% = 3.95 mm and b = 19.5 mm .
127

taking note of the fact that even though the gear pair rotates in windage conditions, the
bearings are lubricated separately to function properly. As seen in Figure 4.9(a), the
predictions match well with the measured Pw values, with the maximum deviation
between measurements and prediction being well within 0.2 kW over the operating speed
range, with the maximum deviation occurring at the highest operating speed value of
= 10, 000 rpm. Deviations at higher values are potentially caused by the simplicity
of bearing power loss formula [4.4] and as well as some of the critical assumptions of the
windage model.
The predicted and measured [4.2] Pw values are compared in Figure 4.9(b) for the
23-teeth gear pair having m% = 3.95 mm and b = 19.5 mm. The agreement between the
two is again acceptable and mostly within the bounds of experimental scatter, with
maximum deviations being well within 0.2 kW. The comparison provided in Figures
4.9(a) and 4.9(b) for the two different gear designs demonstrates the accuracy of the
windage model.

In spite of the fact that the proposed model was a reduced-order

analytical model that employed many simplifying assumptions, the observed level of
accuracy is rather encouraging, given the fact that computational effort required is
negligible.
While gear pair spin loss experiments like the one presented in ref. [4.2] are
sparse, literature includes several experimental studies on windage losses of a single gear.
In such cases, the windage losses are limited to air drag as no pocketing losses can occur.
In one such study, Dawson [4.3] performed a large number of spur and helical gear
windage loss measurements. These measurements were curve-fit to derive an empirical
128

power loss formula for a single gear operating in free air that takes into account the
operating speed , gear outside radius ro and air properties (density and kinematic
viscosity k ) as [4.3]

2.85 4.7
Pw = a0.15
ro .
k

(4.2)

where a is an empirical constant. Meanwhile, Eq. (3.19) from Chapter 3 computes the
drag losses on faces of a gear in the turbulent regime as

2.86 4.72
Pw = 0.0250.14
ro .
k

(4.3)

This empirical formula, Eq. 4.2, of Dawson is very much identical to the
analytical formula of Eq. (4.3) for the face drag power loss in turbulent regime, not only
in terms of the parameters involved but also in terms of the values of the exponents.
Dawsons empirical formula and Eq. (4.3) from the windage power loss model both have
, k , i and roi as the only parameters influencing drag power loss. In addition,
Dawsons exponents for , k , i and roi are 1.0, 0.15, 2.85, and 4.7, as seen in Eq.
(4.2) compared to 1.0, 0.14, 2.86 and 4.72 in Eq. (4.3), respectively. This indicates that
the proposed physics-based drag windage power loss formulation is indeed valid within
the range of experimental scatter as seen through Dawsons measurements.

129

4.4 Concluding Remarks


In this chapter, oil churning and windage power loss models proposed in earlier
chapters were used to simulate the experiments of two companion studies [4.1, 4.2] on
load-independent spin power losses of spur gear pairs operating under dip- and jetlubricated conditions. These experiments were performed over a wide range of operating
speed, temperature, oil levels and key gear design parameters, to allow for quantification
of their influence on spin power losses. Through comparisons to these measurements, the
physics-based oil churning and windage power loss models were shown to be capable of
predicting the measured spin power loss values under both oil churning and windage
conditions, within a reasonable degree of accuracy. It was also shown that the models
predict the parameter sensitivities and experimental trends accurately. In addition, a
close agreement between the proposed windage power loss model formulations and a
published empirical formula [4.3] was reported for further validation of the models. In
view of these comparisons, it can be concluded that the models are realistically accurate,
and hence, can be used for larger scale transmission power loss simulations.

130

References for Chapter 4


[4.1]

Moorhead, M, D., Experimental investigation of spur gear efficiency and the


development of a helical gear efficiency test machine, MS thesis, The Ohio State
University, Columbus, Ohio, 2007.

[4.2]

Petry-Johnson, T. T., Kahraman, A., Anderson, N.E., and Chase, D. R., 2008, An
Experimental Investigation of Spur Gear Efficiency, ASME Journal of
Mechanical Design, 130 (6), 062601, 10 pages.

[4.3]

Dawson, P. H., 1984, Windage Loss in Larger High-Speed Gears, Proceedings


of the Institution of Mechanical Engineers, Part A: Power and Process
Engineering, 198(1), 5159.

[4.4]

Harris, A. T., 2001, Rolling Bearing Analysis, Fourth Edition, Wiley & Sons, Inc.,
New York.

[4.5]

Petry-Johnson, T., Experimental Investigation of Spur Gear Efficiency, MS


thesis, The Ohio State University, Columbus, Ohio, 2007.

131

CHAPTER 5

APPLICATION OF OIL CHURNING POWER LOSS MODEL TO AN


AUTOMOTIVE TRANSMISSION

5.1 Introduction
With the oil churning power loss model of a gear pair, which was proposed in
Chapter 2, having been validated through comparisons to experiments in Chapter 4, the
last remaining objective of this dissertation is the expansion of the formulations to predict
spin power losses of an entire transmission. Given the motivations listed in Chapter 1 in
regards to fuel economy and emissions goals imposed on automotive drive trains and
transmissions, achieving a spin loss model would be valuable in quantifying and
minimizing such losses.
The literature on power losses of geared transmissions is very sparse, with most
studies adapting existing empirical and dimensional-analysis based formulations to
predict friction and spin losses in gear pairs in transmissions. Changenet and Velex [5.1]
proposed a power loss model for a 6-speed manual transmission using thermal networks.
The model predicts temperature distribution and efficiency by considering power loss due
to rolling and sliding friction, rolling bearing elements and oil shearing taking place in

132

the synchronizers. Heingartner and Mba [5.2] proposed a methodology for estimating
power losses in high-speed gear units by combining an empirical power loss expression
for oil churning and windage [5.3], along with a rolling and sliding friction loss model
[5.4]. The predictions of the combined empirical models for friction and viscous losses
in the gear unit were compared with experiments conducted by the same authors on an
example single-stage double-helical speed-increasing unit.

This study lead them to

conclude that increasing the load at a constant speed increased the sliding losses and
slightly decreased the rolling friction losses, while the windage losses remained constant.
While their study is a useful addition to the literature on transmission power loss, no
insights were thrown on the physics behind the mechanism of transmission power losses.
Changenet and Velex [5.5] continued on the same vein of thought by providing for an
experimental versus theoretical comparison of power losses in a six-speed manual
transmission.

The theoretical estimates were again carried out through empirical

formulations for friction and oil churning power losses [5.5], and therefore, they were not
physics-based. From their comparisons, they inferred the transmission efficiency to be
higher at higher operating temperature and higher load values. Handschuh and Kilmain
[5.6] carried out experiments on characterizing the efficiency of high-speed helical gear
trains that find extensive use in tilt-rotor aircrafts. The results of the experimental study
were compared to empirical estimates for power losses due to friction [5.4] and viscous
dissipation [5.7] in the gear meshes and bearings. They concluded that power losses
increased with speed, with torque transmitted by the gear train having a negligible effect
on the power losses. van Dongen [5.8] quantified the influence of speed and load on
transmission efficiency in manual and automatic transmissions through dimensional
133

analysis, leading to the conclusion that transmission efficiency increased as load


increased, with oil churning acting as the dominant source of power loss in
transmnissions. In general, the complexity of the environment inside the gearbox as well
as difficulties in measuring gear tooth temperature has made the study of transmission
power losses very challenging.
A critical review of the above studies only goes on to show that transmission
power losses were tackled from the point of view of approximate analyses, with major
limitations existing in the application of such empirical formulae in estimating power
losses in transmissions, as experimental conditions do not remain consistent.

Also,

constraints on computational time and effort prevent the seeking of solutions through a
CFD-based methodology.

Hence, applicability of such models to larger multi-gear

transmissions has been cumbersome.


With the above thoughts in mind, this chapter will focus on the prediction of the
total transmission power loss of a manual transmission, which is given by the sum of the
power losses at the gear mesh due to friction, and the spin power losses, as well as the
power losses due to mechanical and viscous dissipation in bearings. As the centerpiece of
this work, the fluid mechanics-based model of oil churning losses of a gear pair as
formulated in Chapter 2 will be utilized here and applied to the transmission as a whole.
This model will then be combined with a transmission mechanical power loss model,
based on a gear pair friction power loss model [5.9] and bearing power loss models [5.10]
to predict the total power loss of a manual transmission. The transmission power loss
predictions will be validated through comparisons to measurements [5.11] from an
134

example 6-speed manual transmission, preformed under tightly controlled test conditions.
Application of the fluid-dynamics based oil churning power loss model and also the
mechanical power loss model of Xu et al [5.9] will provide insight into the physics behind
the transmission power losses.
In general, the total transmission power loss PT is comprised of two groups of
losses, as explained in the introductory chapter: (i) load-dependent (friction induced)
mechanical power losses Pm and (ii) load-independent (spin) losses Ps , i.e. PT = Pm + Ps .
Friction-induced losses at the loaded gear meshes, Pmg , and at the bearings, Pmb , define
the load-dependent mechanical power losses. The total mechanical loss is the sum of
losses from all loaded gear meshes and bearings: Pm = Pmg + Pmb .

Load-independent (spin) losses are caused by a host of factors, as detailed in


earlier chapters, including viscous dissipation of bearings, Psb , churning of oil and
windage by gears, Psg , and are given as Ps = Psg + Psb ; the physics of the loadindependent power losses has been laid out in Chapters 2 and 3 of this dissertation, in
sufficient detail. Other secondary losses in components such as synchronizers and oil
seals can also be added to this total. As stated in preceding chapters, while losses from
these two groups might be comparable under high-load and low-speed conditions, the spin
losses tend to dominate over the load-dependent power losses at higher speeds and lighter
loads.

Section 5.2 will provide an introduction to the methodology of manual

transmission power loss prediction, with sub-sections outlining gear pair spin and
mechanical power loss models in brief. Finally, the transmission power loss model will be
135

applied to an actual manual transmission and compared to transmission power loss


measurements from a companion study on an example six-speed manual transmission for
validation purposes.

5.2 Manual Transmission Power Loss Prediction Methodology


The transmission power loss prediction methodology comprises of three stages of
power loss computations as illustrated in the flowchart of Figure 5.1. The first task in this
methodology is to determine (i) the load transmitted by each gear pair, (ii) forces acting on
each bearing (including preloads), (iii) speed of each gear, and (iv) oil level at each gear
location. These calculations, represented by Kinematics and Power Flow in Figure 5.1,
are repeated at each duty cycle increment j, characterized by the transmission input speed

j , input torque T j , gear state g j , and a user defined operating temperature j and
static oil level (or oil volume). These operating conditions and corresponding bearing,
gear and transmission geometric data are then used to compute (1) load-independent spin
losses of losses of each rotating gear pair and bearings and (2) load-dependent
(mechanical) losses of every load carrying gear mesh and bearings. All these components
of power losses are added to represent the losses at a transmission cycle increment j and
the same computations are repeated for all discrete increments to determine the variation
of the power losses with a particular fuel economy cycle.

136

Lubricant
Conditions
j , Oil Level

Gear Parameters
Bearing Parameters
Transmission Geometry

Duty Cycle
Tj , j , g j

Kinematics &
Powerflow Analysis

Bearing Mech.
Loss Model

Gear Mech.
Loss Model

Gear Spin
Loss Model

Pmb

Pmg

Psg

Pm

Bearing Spin
Loss Model

Psb

Ps

PT
Total Transmission
Power Loss

Figure 5.1 Flowchart of the transmission power loss computation methodology.

137

5.2.1 Gear Pair Spin Power Loss Model.


The gear pair spin power loss model employed in the computation of loadindependent power losses in the transmission has been described in detail in Chapter 2.
Accordingly, the total gear pair spin power loss will be given as the sum of the drag loss
on the periphery of the gears, the drag loss on the faces of the gears, power loss due to root
filling in the cavity between adjacent gear teeth, and finally, the power loss due to
pocketing/squeezing of the lubricant in the meshing zone.

With drag loss on the

periphery, faces and power loss due to root filling of gear i denoted as Pdpi , Pdfi and Prfi ,
respectively, and pocketing power loss in the meshing zone given as Pp , and expressions
for the same given by Eqs. (2.7b), (2.13), (2.23) and 2.42(b), respectively, for the laminar
regime (with Eq. (2.17) replacing Eq. (2.13) if flow along the faces is turbulent), the total
gear pair spin power loss is given as

Psg = Pdp1 + Pdp 2 + Pdf 1 + Pdf 2 + Prf 1 + Prf 2 + Pp .

(5.1)

5.2.2 Gear Pair Mechanical Power Loss Model.


As illustrated in the flowchart of Figure 5.1, the gear mesh mechanical power loss
model combines a gear load distribution model, a gear contact friction model and a gear
mechanical efficiency formulation [5.9]. The gear load distribution model predicts the
load and contact pressure distribution at every contact point over the tooth surface of the
gear, as the mesh position is incremented. The corresponding load distribution along with
138

geometric parameters is input to the gear contact friction model, which determines the
distribution of friction coefficient along the contacting surfaces. This friction coefficient
value is then input to the mechanical efficiency formulation to determine the instantaneous
power loss at an incremental rotational position m ( m [1, M ] ). The same procedure is
repeated at other discrete rotational positions until m = M , when a complete gear mesh
cycle of rotation is achieved. The instantaneous mechanical efficiency values over the
entire mesh cycle are then averaged to obtain the average mechanical power loss of the
gear pair at the given instantaneous duty cycle condition.
Xu et al [5.9] used a thermal EHL model as the basis to predict distribution along
the gear contacts. In order to overcome the high computational demand of the EHL
model, they performed a large number of EHL analyses for the lubricant considered,
within typical ranges of all key contact parameters including surface characteristics,
operating conditions and contact pressure distribution. Then a multiple linear regression
analysis of the analyzed EHL results was performed to reduce these simulations into a
single friction coefficient formula given as [5.9]

b1 +b4 SR Ph log10 ( 0 )+b5 exp( SR Ph log10 0 )+b9e S Pb2 SR b3 Veb6 b7 Rb8


0
h
= e

(5.2)

This formula contains most of the key parameters including the absolute viscosity of the
lubricant 0 (in cPs), the RMS composite surface roughness S (in m), the effective
radius of curvature R (in meters) and the contact pressure Ph (in Pa). SR is the sliding
ratio SR = 2 (u1 u2 ) (u1 + u2 ) where u1 and u2 are the velocities of the contacting

139

surfaces in the sliding direction.

In Eq. (5.2), bi (i = 1, 2,..., 9) are constant coefficients

that are established for the manual transmission fluid of interest. This formula was shown
to agree well with both actual EHL analyses and traction measurements [5.9]. After
finding the friction coefficient ( m) ( z , ) at each contact segment centered by the
coordinates ( z , ) at each incremental rotational position m = [1, M ] , the friction force is
calculated as [5.9]

F ( m ) ( z , ) = ( m ) ( z , )W ( m ) ( z , ) ,

(5.3a)

where W ( m) ( z , ) is the normal force at the center of the contact segment. After the
friction forces at each loaded contact point q [1, Q ] are computed, the instantaneous gear
mesh mechanical power is obtained as [5.9]

( m)
Pmg
= Pin F ( m) (u1 u2 ) .

(5.3b)

q =1

Finally, average value of mechanical power loss over a mesh cycle is computed as [5.9]

Pmg =

1
M

Pmg(m) .

(5.3c)

m =1

140

5.2.3 Bearing Power Loss Model


The total bearing power loss is given by the sum of the bearing mechanical power
loss and bearing viscous (spin) power loss. For the computation of the bearing mechanical
power losses, the axial and radial components of forces acting on bearings placed at either
ends of a shaft s are computed at each duty cycle increment, for a given input torque, input
speed and average bulk temperature value. The preload force, if any, is included in the
computation of the axial component of the bearing force. Once the effective bearing force
is obtained from force and moment balance, the friction moment on a bearing i located on
shaft s is computed as [5.10]

(i , s )
M mb
= F d ,

(5.4a)

where is a friction factor, d = 0.5(d o + di ) is the mean bearing diameter (in mm), and
F is the effective bearing force. The value of F is taken to be the greater of the radial

bearing force or the combined product of

4 cot
5

times the axial bearing force, where is

the bearing contact angle. With a rotational shaft speed of s , the power loss due to
(i ,s )
(i ,s )
= M bm
s . Similarly, the
friction at i bearings of a given shaft s is computed as Pbm

spin torque loss for a bearing i on a shaft s is given by: [5.10]

(i , s )
M sb

(107 ) f L k 2 / 3 2 / 3d m3

=
160(107 ) f L d m3

k > 2000
k < 2000
141

(5.4b)

where f fl is a bearing viscous friction constant, ns is the shaft speed in rpm and k is
the kinematic viscosity of the lubricant in mm2/s. With this, the spin power loss of the
(i ,s )
s .
same bearing is found as Psb(i ,s ) = M sb

5.3 Application of the Power Loss Models to an Actual Transmission


In this section, the manual transmission power loss methodology shown in the
flowchart of Figure 5.1 will be applied to a production manual transmission. This 6-speed
manual transmission shown in Figure 5.2 will be considered as the example transmission.
This transmission consists of four shafts, input, upper output, lower output and
differential. All the gears on the input shaft are fixed to it so that they rotate at the input
speed. As all speed gears on both output shafts are engaged with their counterparts on the
input shaft at all instances, they also rotate at speeds proportional to the input speed
according to their gear ratios. The speeds of the final drive gears, on the other hand are
found from the speed gear and final drive ratios. With this and the duty cycle data ( T j ,

j , g j ( g j [1, 6] ), rotational speeds of all of the gears are calculated in the Kinematics
and Powerflow routine. Likewise, individual immersion depths of each gear pair are
determined based on the volume of oil in the transmission. In addition, bulk transmission
temperature j at a discrete time instant j ( j [1, J ] ) is input to the transmission power
loss model.
Figure 5.3 shows the orientation of a total of 9 gear pairs (15 gears) of the 6-speed
example manual transmission for given arbitrary static oil level. With this oil level,
142

Figure 5.2 Example system: A 6-speed manual transmission.

143

Figure 5.3 Rotating gears on the planes of (a) 6th gear pair, (b) 4th gear pair, (c) 3rd and
5th gear pairs, (d) 2nd gear pair, (e) 1st and reverse gear pairs, and (f) final
drive gear pair.
144

immersion parameter h of each gear is determined and drag spin loss calculation are
carried out using the model of Section 5.2.1 for all 15 gears. Likewise, pocketing losses
for each of the 9 gear meshes are also computed using the pocketing loss model of the
same section. If a gear is not immersed in oil, its oil churning drag loss is assumed to be
zero and the windage loss is computed instead by considering the properties of air
replacing that of oil. Similarly, the pocketing loss of any gear mesh above the static oil
level is assumed to be negligible. For the 6-speed manual transmission, the total gear spin
loss at a given duty cycle instance j is then found as

15

i =1

r =1

Psg = [ Pd ]i + Pp .
r

(5.5a)

As each shaft in this transmission is supported by a pair of tapered roller bearings, the spin
power loss due to bearings is given as

Psb = Psb(i , s ) .

(5.5b)

s =1 i =1

With this, the total spin power loss of the transmission due to gears and bearings is
Ps = Psm + Psb .
Next, following the mechanical loss branch of Figure 5.1 and employing the model
presented in Section 5.2.2, the mechanical gear mesh losses are computed.

These

computations are done only for the gear pairs that carry load. In any given forward gear
145

state, g j [1, 6] , this transmission has only two loaded gear pairs. For instance, the speed
gear pair (sg) in Figure 5.3(e) and the lower final drive (fd) gear pair of Figure 5.3(f) are
the only load carrying gear meshes that must be considered in mechanical loss
computations. Accordingly, total mechanical gear mesh loss of the transmission at the
same duty cycle instance is given as

( sg )
( fd )
Pmg = Pmg
+ Pmg
,

(5.6a)

Likewise, for each bearing with its load computed (including preload), the
mechanical power loss from all the eight bearings of the 6-speed sample manual
transmission is given as

(i ,s )
Pmb = Pmb
,

(5.6b)

s =1 i =1

and the total mechanical power loss of the transmission due to gears and bearings is found
as Pm = Pmg + Pmb .

Finally, the total power loss of the transmission is given as

PT = Ps + Pm .

146

5.4 Transmission Power Loss Experiments and Model Validation


Chapter 4 details the validation of the gear pair spin power loss model. Hence, the
validation activity in this chapter will be performed at the transmission level. For this
purpose, an experimental setup was designed to measure the mechanical and spin power
losses of the 6-speed manual transmission over a wide range of input torque, operating
speed and bulk transmission temperatures [5.11]. In this section, a brief description of the
experimental procedure will be provided and direct comparisons between the power loss
measurements and predictions will be made to assess the accuracy of the proposed model.

5.4.1 Test Set-up and Testing Methodology


Figure 5.4 shows the dynamometer setup that was designed and implemented for
measurement of power losses of the 6-speed manual transmission through tests carried out
by Szweda [5.11]. A more detailed description of the experimental setup is provided in
ref. [5.11]. Describing only the necessary details, a 110 kW DC drive motor is connected
to the example manual transmission through a slender shaft, a pair of flexible couplings
and a precision torque-meter at the input side. A support bracket and an adapter plate are
used to hold the transmission in place at its intended position. On the output side, a
precision torque-meter is connected to the output shaft by a flexible coupling. On the
other side of the output torque-meter, a long shaft and another flexible coupling is used to
connect to a 1.6 ratio speed-increaser that is connected to an eddy-current load brake
dynamometer (400 kW). A thermocouple inserted from the bottom drain hole measures
the bulk transmission temperature.

147

Figure 5.4 Experimental test setup for measuring transmission power losses of the
example 6-speed transmission.
Safety guards are removed for
demonstration purposes.

148

Measurements were repeated at various input speeds ( j = in ) and gear stages,

g j , for both the loaded and unloaded test cases. The input speed was maintained within
20 rpm and the input torque in the loaded test case was maintained within 2 Nm.
First, unloaded tests were conducted with output shaft disconnected. The torque, Tin ,
measured by the input-side torque-meter was used to compute the transmission spin power
loss Ps = jTin . Repeating the same test with the output side connected and applying a
nominal input torque value of T j , the power loss of the transmission was found as

PT = j (Tin Tout j ) where j is the input-to-output speed ratio and Tout is the torque
value measured by the output-side torque-meter.

5.4.2 Model Validation


Figure 5.5 provides for comparison between the measured and predicted
dimensionless transmission mechanical power loss values, Pm = Pm Pref , where Pm is the
transmission mechanical power loss value and Pref is a reference power loss value for the
example 6-speed manual transmission. The comparisons are done an input operating
speed range of 1,000 to 4,000 rpm and for an input torque of 100 Nm at 80o C with the
production oil volume. The composite root-mean-square surface roughness of each gear
pair was measured to be S = 0.56 m. As is apparent from the figure, the predicted and
measured dimensionless transmission mechanical power loss values show good agreement
over the entire range of operating conditions and gear stages, with maximum deviations

149

0.5

Experiment

(a) 1000 rpm

(b) 2000 rpm

(c) 3000 rpm

(d) 4000 rpm

Prediction

0.4

Mechanical Power Loss Pm

0.3
0.2

0.1
0.0
0.5
0.4

0.3
0.2

0.1
0.0
1

Gear Range

Figure 5.5

Comparison of measured [5.11] and predicted transmission mechanical


power loss values.

150

between the measured and predicted Pm values being well within 0.1 kW. This suggests
that the transmission mechanical power loss model elucidated in Section 5.3 is sufficiently
accurate for studies to characterize friction losses in gear trains and transmissions and can
also be used to study methods for reduction of the load-dependent friction power losses.
Figure 5.6 provides for comparisons between the measured and predicted
dimensionless spin power loss values, Ps = Ps Pref , where Ps is the transmission spin
power loss and Pref is the same reference power loss value, for gear stages 2 to 6 and for
a range of operating speed, in = 1,000 to 4,000 rpm. Here, the same volume of oil at
80o C was used for these tests. As shown in Figure 5.6, the predicted and measured spin
power loss values agree quite well over the entire operating range, with the predictions
remaining consistently lower than the measured spin power loss values, though following
the same trend as the measurements. The maximum deviation between the predicted and
measured Ps values are within about 0.3 kW. Several factors might be the reason for this
difference between the measurements and predictions, the most important of them being
the fact that the spin loss model captures only the gear and bearing related losses.
Inclusion of the other sources of spin power losses such as synchronizers and seals should
bring the predictions even closer to the measurements. Also, the oil churning model uses
static oil levels to compute the immersion depths of gears in churning. Under dynamic
conditions, effective oil levels should be expected to be higher, especially for the upper
chamber of the transmission, increasing the predicted losses further. It is also important to
note that the bearing viscous power loss model used in this study is rather simplistic as it
151

2.0

Experiment

(a) 2nd Gear

(b) 3rd Gear

(c) 4th Gear

(d) 5th Gear

Prediction
1.6

1.2

Spin Power Loss Ps

0.8

0.4

0.0
2.0

1.6

1.2

0.8

0.4

0.0
0

1000

2000

2.0

3000

4000

Experiment

1000

2000

3000

4000

(e) 6th Gear

Spin Power Loss Ps

Prediction
1.6

1.2

0.8

0.4

0.0
0

1000

2000

3000

4000

Input Speed [rpm]

Figure 5.6 Comparison of measured [5.11] and predicted transmission spin power loss
values: (a) 2nd gear, (b) 3rd gear, (c) 4th gear, (d) 5th gear, and (e) 6th gear.
152

cannot account for several key parameters, most importantly the oil immersion depth.
Therefore, bearing viscous power loss predictions are expected to be less accurate. On the
whole, these comparisons indicate that the transmission spin power loss model formulated
and developed is indeed effective in predicting the spin power loss of gear pairs under
different configurations in a transmission in a computationally effective way, thus
providing confidence to end users.

5.5 Concluding Remarks


In this chapter, a physics-based methodology was proposed to predict the
mechanical and spin power losses of geared transmissions. The transmission power loss
model incorporates a transmission spin power loss model and a transmission mechanical
power loss model in conjunction with bearing mechanical and viscous power loss models
in order to predict the load-dependent and load-independent power losses of all the gear
pairs in a transmission. Comparisons of the predictions to tightly controlled transmission
power loss measurements demonstrate good correlation over wide ranges of operating
speed, input torque and gear range, indicating that the model can be used for design and
product improvement activities.

153

References for Chapter 5


[5.1]

Changenet, C. and Velex, P., 2005, Power Loss Predictions in a 6-speed Manual
Gearbox Using Thermal Networks, ASME Design Engineering Technical

Conferences and Computers and Information in Engineering Conference,


DETC2005-84290, Long Beach, CA..
[5.2]

Heingartner, P., and Mba, D., 2003, Determining Power Losses in The Helical
Gear Mesh; Case Study, Proceeding of DETC3, ASME 2003 Design

Engineering Technical Conferences and Computers and Information in


Engineering Conference, Chicago, I
[5.3]

Luke, P., and Olver, A., 1999, A Study of Churning Losses in Dip-Lubricated
Spur Gears, Proc. Inst. Mech. Eng.: J. Aerospace Eng., Part G, 213, 337346

[5.4]

Anderson, N. E., and Loewenthal, S. H., 1982, Design of Spur Gears for
Improved Efficiency, ASME Journal of Mechanical Design, 104, pp. 767774..

[5.5]

Changenet, C., and Velex, P., 2007, A Model for the Prediction of Churning
Losses in Geared Transmissions Preliminary Results, ASME Journal of

Mechanical Design, 129(1), pp. 128-133.


[5.6]

Handschuh, R. F., Kilmain, C. J., 2003, Efficiency of High-Speed Helical Gear


Trains, 59th Annual Forum and Technology Display sponsored by the American
Helicopter Society, Phoenix, AZ.

154

[5.7]

Dawson, P. H., 1984, Windage Loss in Larger High-Speed Gears, Proceedings

of the Institution of Mechanical Engineers, Part A: Power and Process


Engineering, 198(1), 5159.
[5.8]

van Dongen, L. A. M., 1982, Efficiency Characteristics of Manual and


Automatic Passenger Car Transaxles, SAE Technical Paper Series, 820741,
Warrendale, PA: Society of Automotive Engineers.

[5.9]

Xu, H., Kahraman, A., Anderson, N.E., Maddock, D.G., 2007, Prediction of
Mechanical Efficiency of Parallel-Axis Gear pairs, ASME Journal of Mechanical

Design, 129, 58-68.


[5.10] Harris, A. T., 2001, Rolling Bearing Analysis, Fourth Edition, Wiley & Sons, Inc.,
New York.
[5.11] Szweda, T, A., An Experimental Study of Power Loss of an Automotive Manual
Transmission, MS thesis, The Ohio State University, Columbus, Ohio, 2008.

155

CHAPTER 6

CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE WORK

6.1 Summary
This dissertation lays the framework for a physics-based fluid mechanics model to
predict oil churning and windage power losses in gear pairs. The oil churning and
windage power loss models, both developed for the case of a spur gear pair, have been
validated through comparisons to power loss measurements conducted on a family of
unity-ratio spur gear pairs, under dip- and jet-lubricated conditions. Validations were
carried out over a wide range of operating conditions, different oil immersion depths,
lubrication conditions and through variations in gear geometry parameters, to enhance the
fidelity of the proposed models. The influence of key system parameters were also
quantified through detailed example analyses of the respective power loss models.
Further, the oil churning power loss model, in conjunction with a mechanical
power loss model and existing bearing power loss models, has been generalized, in the
form of a transmission power loss model, to predict the total power loss in automotive
manual transmissions.

Power loss measurements on a sample six-speed manual

transmission corroborate the predictions of the proposed transmission power loss model.
156

The power loss computation methodologies proposed in this dissertation have the
potential to be generalized so as to form the foundation for future formulations that can
handle other types of complex gear drives, beyond just spur and helical. It cannot be
more highly stressed that in lieu of the lack of physics-based models that exist in
transmission power loss literature, the models developed in this dissertation, with all their
limitations, present a detailed first-stab towards the prediction of power losses in gears
without compromising on computational efficiency in providing a clear window into the
mechanism of such power losses in geared systems.

6.2 Conclusions and Contributions


The comprehensive modeling efforts to predict oil churning and windage power
losses in gear pairs, backed by experimental validations on both the gear pair and the
transmission levels, have resulted in the following general conclusions:

Predictions of the oil churning power loss model compare reasonably well with
dip-lubricated spin loss measurements. Concurrently, predictions of the windage
power loss model show a good correlation with jet-lubricated experiments. This
proves the effectiveness of both of the model prediction methodologies. The
transmission power loss model also compares well with measurements from a
sample six-speed manual transmission, which makes the transmission power loss
model suitable as a design tool.

157

Oil immersion depth has maximal influence on power loss due to oil churning,
with a substantial increase seen in oil churning losses as the oil immersion depth
increases. In the case of transmission spin losses, the influence of oil immersion
depth is also as significant, with greater spin losses predicted at higher oil fill
levels.

Power loss due to oil churning and windage decrease with an increase in
operating temperature, regardless of operating speed. Correspondingly, oil
churning and windage power losses are influenced heavily by operating speed,
registering substantial increase with an increase in gear rotational speed. In the
case of power losses in a transmission, the influence of operating speed is more
pronounced at higher gear stages.

Face width, gear module and outside diameter influence the total spin power loss
to a much lesser extent than lubricant parameters and operating conditions, with
comparatively minor increases seen in the model predictions with an increase in
gear geometry parameters. Surface roughness plays no part in influencing either
the oil churning or windage modes of power losses.

Power loss due to oil churning is several times higher than power losses loss due
to windage, under similar operating conditions.

Also, drag and pocketing

components of load-independent power losses overshadow the power loss due to


root filling.

158

This dissertation research provides a fundamental understanding of the


mechanisms of spin power losses of gear pairs as well as of larger transmission systems.
The following can be listed as the main contributions of this study to the sate-of-the-art.

This study has provided the first physics-based analytical model for oil churning
and windage losses of a spur gear pair. As previous efforts relied on empirical or
CFD type approaches, their capabilities were rather limited, both computationally
and methodologically. The analytical and closed-form nature of the proposed oil
churning and windage power loss models makes possible for computationally
effective power loss estimations under various operating conditions, even during
the early stages of product design, allowing for discrimination of various design
concepts based on their efficiency attributes.

The spin power loss model developed in this study represents a general
formulation, which has the potential to form a foundation for similar such models
for hypoid gears in automotive axles (a significant source of power losses) and
planetary gear sets in automatic transmissions.

This study also constitutes the first theoretical model for predicting both loaddependent and load-independent power losses of multi-mesh gear systems such as
manual transmissions and gearboxes. While being limited in its contributions to
the fundamental aspect of this topic, the proposed model has the potential to
impact upon the intended goals of this study in terms of fuel economy of vehicles
with manual transmissions. The model proposed will include all key system

159

parameters so that cost effective strategies to lower transmission power loss can
be developed.

The proposed gear pair and transmission power loss models will also distinguish
themselves as the first validated power loss models.

Comparisons to

measurements from tightly controlled gear pair experiments as well as


experiments conducted on a six-speed manual transmission make for a complete
multi-parameter validation of the models at both gear pair and transmission levels.
This provides added confidence to end users.

6.3 Recommendations for Future Work


Although the proposed models are exhaustive in their approach, they are by no
means capable of addressing all the functional issues that come into play when geared
systems are looked into in their full glory. As the topic of efficiency of automotive
applications is becoming more crucial, especially in high power applications, and with
increasing demands placed on conservation of fossil-based resources, it is vital to expand
on some of the stated objectives of this dissertation, as a means to enhance the accuracy
of the current model predictions. The following are some recommendations for future
work:

Effect of enclosures: Drag power losses in the case of oil churning and windage
modes of power loss did not include the effects of enclosures. The presence of
peripheral shrouds, baffles and flanges will certainly change the nature of the
160

external flow or boundary layer between the enclosure and the gear surfaces. As
such, the current formulations are valid for as long as the distance between the
enclosures and the gear surfaces is greater than the combined thickness of
boundary layers from both the surfaces. Future modifications of the formulations
should look into the effects of peripheral clearance on churning and windage
losses. This can have potential impact on noise and vibrations induced through
pocketing power losses as well, with an increase in the intensity of noise produced
upon impact of the ejected jet of lubricant from the meshing zone as enclosures
wrap gear pairs more tightly.

Presence of form drag: The drag formulations detailed in Chapters 2 and 3 take
into account only the skin friction drag while calculating the total drag loss on the
periphery and faces of the gears, before boundary layer separation has taken
place.

The effects of pressure or form drag are not included based on the

assumption that the wake region formed on boundary layer separation is


negligible. This has to be looked into at greater depth due to the fact that form
drag can potentially increase churning power loss at low viscosity values and
higher speeds of operation. The inclusion of form drag will need more details on
the nature of flow separation and adverse pressure gradient leading to separation.

Dynamic oil levels: All through the power loss formulations, it has been assumed
that oil immersion depths are quasi-static in nature, with a broad based
supposition that a reduction in oil level in one part of the chamber encasing the
gear pairs will result in an increase in oil level at another part of the chamber,
161

leading to an averaging effect, which aids in the assumption. In reality, oil levels
are highly dynamic in nature, with oil swirling all around the gear faces and
periphery, making it very difficult to peg onto a particular oil level value for
computation purposes. The inclusion of dynamic oil levels, through means of a
fundamental relationship between oil level, operating speed, enclosure/gear size
and lubricant conditions, can greatly aid in improving the accuracy of power loss
predictions.

Improvements to the pocketing power loss formulation: It has been consistently


shown that pocketing power loss is the major component of the total gear pair
spin power loss. Future revisions to the pocketing component of the spin power
loss model must include the effects of a moving control volume approach, and the
motion of the control surface. Further more, flow through the pockets is threedimensional, with a flow factor in the current formulation taking this into account.
A more detailed three-dimensional analysis can account for such assumptions of
the current work. Also, during the modeling of pocketing power loss, it has been
assumed that there exists no pressure differential between opposing control
volumes thereby leading to the consideration of the control volumes as separate
entities in time.

A renewed and in depth approach will consider the

communication between opposing control volumes as a continua, which will


augment the mass flow rate of the fluid ejected through the end flow areas.

Contribution of helix angle: The spin power loss model has been developed for
the simplest case of spur gears. In most transmission applications, the gears used
162

are helical in nature, for obtaining higher speeds and smoother progression of
mating, leading to a reduction in operating noise. When the spin power loss
model is applied to the scenario of a pair of helical gears, changes in the flow
areas due to helix angle has to be considered when applying the pocketing power
loss formulation. On the whole, power losses are expected to be lesser for helical
gears than in the case of spur gears; the formulations have to be subtended to take
this into account.

Surface roughness of gears: In the turbulent regime, while modeling drag losses
on the faces and periphery, surface roughness might potentially change the
behavior of drag forces, leading to earlier flow separation by enhancing the
intense mixing taking place in the boundary layer. This effect can be considered
while modifications to the drag losses are being made. Also, at lower speeds of
rotation, the boundary layer characteristics might change from that of a growth
around a disk. Future revisions to the drag loss formulations should take this into
account.

Effects of air-ingestion: Air-ingestion, which can take place for medium oil
levels and at high speeds, will change the atmosphere prevailing in the meshing
zone, thereby changing the nature of dependence of density on temperature and
pressure.

Hence, the analysis itself will then borders on a two-phase flow

approach, increasing the complexity of the problem but also making it a more
complete approach through such an inclusion. Partial pressures, densities and
viscosities can be utilized in such a case, depending on the aeration ratio.
163

Effect of rotational direction: Experimental evidence has shown that a change in


the direction of gear pair rotation from up-in-mesh to down-in-mesh reduces the
total spin power losses of a gear pair. A down-in-mesh rotational arrangement
will result in lesser amount of lubricant squeezed through the meshing zone, but
greater amounts of air ingestion, leading to a methodology, which will need a
combined compressible-incompressible flow approach. This approach, though
very obvious, is not easy, as it is extremely hard to deduce the fraction of air to oil
ratio in a gear mesh (aeration ratio), at a given rotational position. This works in
superposition with the effects of air-ingestion as explained in the previous point.

Application to hypoid and planetary gear sets:

Subtle changes in the

formulations, in terms of tooth cavity volumes and immersion areas have to be


made to accommodate other gear geometries, such as cross-axis gears and
planetary gear sets. Future formulations will focus on such geometrical aspects.
On a final note, while the intended goal of this dissertation is to come up with a
physics-based methodology for predicting spin power losses in gear pairs and
transmissions, along the way, broad-based assumptions have been impressed upon the
formulation to eschew computational difficulties, with proper justifications. Based on
these assumptions, which certainly do not detract from the ease of use of the proposed
models, it has to be noted that a point-to-point comparison between measurements and
model predictions is certainly not the aim of this dissertation; the main focus of coming
up with such a methodology is to predict the trends of load-independent power losses in
gear pairs from a simplistic but effective engineering viewpoint. In that sense, this
164

dissertation achieves what it set out to accomplish. It is hoped that future users of the
models formulated in this dissertation will understand and appreciate the limitations of
this work in terms of the complexity of the problem handled and the apparent sensitivity
of the power loss predictions to a whole host of parameters other than the most important
ones considered in this work.

165

BIBLIOGRAPHY

Akin, L. S., and Mross, J. J., 1975, Theory for the Effect of Windage on the Lubricant
Flow in the Tooth Spaces of Spur Gears, ASME Journal of Engineering for Industry, 97,
12661273.

Akin, L. S., Townsend, J. P., and Mross, J. J., 1975, Study of lubricant jet flow
phenomenon in spur gears, Journal of Lubrication Technology, 97, 288-295.

Al-Shibl, K., Simmons, K., and Eastwick, C. N., 2007 , Modeling Gear Windage Power
Loss From an Enclosed Spur Gears, Proceedings of the Institution of Mechanical
Engineers, Part A, 221(3), 331341.

Anderson, N. E., and Loewenthal, S. H., 1982, Design of Spur Gears for Improved
Efficiency, ASME Journal of Mechanical Design, 104, pp. 767774.

Ariura, Y., Ueno, T., and Sunamoto, S., 1973, The lubricant churning loss in spur gear
systems, Bulletin of the JSME, 16, 881-890.

Blok, H., 1957, Measurement of Temperature Flashes on Gear Teeth under Extreme
Pressure Conditions, Proceeding of The Institution of Mechanical Engineers, 2, 222235.

Bones, R.J., 1989, Churning Losses of Discs and Gears Running Partially Submerged in
Oil, Proceedings of the ASME 5th International Power Transmission and Gearing
Conference, Chicago, 355-359.

Changenet, C. and Velex, P., 2005, Power Loss Predictions in a 6-speed Manual
Gearbox Using Thermal Networks, ASME Design Engineering Technical Conferences
166

and Computers and Information in Engineering Conference, DETC2005-84290, Long


Beach, CA.

Changenet, C., and Velex, P., 2007, A Model for the Prediction of Churning Losses in
Geared Transmissions Preliminary Results, ASME Journal of Mechanical Design,
129(1), pp. 128-133.

Daily, J. W. and Nece, R. E., 1960, Chamber Dimensional Effects on Induced Flow and
Frictional Resistance of Enclosed Rotating Disks, ASME Journal of Basic Engineering,
82, 217232.

Dawson, P. H., 1984, Windage Loss in Larger High-Speed Gears, Proceedings of the
Institution of Mechanical Engineers, Part A: Power and Process Engineering, 198(1), 51
59.

Diab, Y., Ville, F., and Velex, P., 2006, Investigations on Power Losses in High Speed
Gears, Journal of Engineering Tribology, 220, 191298.

Diab, Y., Ville, F., Houjoh, H., Sainsot, P., and Velex, P., 2005, Experimental and
Numerical Investigations on the Air-Pumping Phenomenon in High-Speed Spur and
Helical Gears, Proceedings of the Institution of Mechanical Engineers, Part C: J.
Mechanical Engineering Science, 219, 785-800.

Eastwick, C. N., and Johnson, G., 2008, Gear Windage: A Review, ASME Journal of
Mechanical Design, 130, 034001, 6 pages.

Handschuh, R. F., Kilmain, C. J., 2003, Efficiency of High-Speed Helical Gear Trains,
59th Annual Forum and Technology Display sponsored by the American Helicopter
Society, Phoenix, AZ.

Harris, A. T., 2001, Rolling Bearing Analysis, Fourth Edition, Wiley & Sons, Inc., New
York.

Heingartner, P., and Mba, D., 2003, Determining Power Losses in The Helical Gear
Mesh; Case Study, Proceeding of DETC3, ASME 2003 Design Engineering Technical
Conferences and Computers and Information in Engineering Conference, Chicago, Il.
167

Hhn, B.R., Michaelis, K., Vllmer, T., 1996, Thermal Rating of Gear Drives: Balance
between Power Loss and Heat Dissipation, AGMA, Fall Technical Meeting, 96FTM8,
pp. 1-12.

Luke, P., and Olver, A., 1999, A Study of Churning Losses in Dip-Lubricated Spur
Gears, Proc. Inst. Mech. Eng.: J. Aerospace Eng., Part G, 213, 337346.

Mann, R.W. and Marston, C.H., 1961, Friction Drag on Bladed disks in Housings as a
Function of Reynolds Number, Axial and Radial Clearance and Blade Aspect Ratio and
Solidity, ASME Journal of Basic Engineering, 83 (4), 719-723.

Martin, K. F., 1978, A Review of Friction Predictions in Gear Teeth, Wear, 49, 201
238.

Moorhead, M, D., Experimental investigation of spur gear efficiency and the


development of a helical gear efficiency test machine, MS thesis, The Ohio State
University, Columbus, Ohio, 2007.

Niemann G., and Lechner G., 1965, The Measurement of Surface Temperature on Gear
Teeth, ASME Journal of Basic Engineering, 11, 641-651.

Pechersky, M. J. and Wittbrodt, M. J., 1989, An analysis of fluid flow between meshing
spur gear teeth, Proceedings of the ASME 5th International Power Transmission and
Gearing Conference, Chicago, 335342.

Petry-Johnson, T. T., Kahraman, A., Anderson, N.E., and Chase, D. R., 2008, An
Experimental Investigation of Spur Gear Efficiency, ASME Journal of Mechanical
Design, 130(6), 062601, 10 pages.

Petry-Johnson, T., Experimental Investigation of Spur Gear Efficiency, MS thesis, The


Ohio State University, Columbus, Ohio, 2007.

Schlichting, H., 1955, Boundary-Layer Theory, McGraw-Hill Book Company, New


York.

168

Soo, S. L., and Princeton, N. J., Laminar Flow Over an Enclosed Rotating Disc,
Transactions of the American Society of Mechanical Engineers, 80, 287-296.

Streeter, L. V., and Wylie, E. B., 1985, Fluid Mechanics, Eighth edition, McGraw-Hill,
New York.

Szweda, T, A., An Experimental Study of Power Loss of an Automotive Manual


Transmission, MS thesis, The Ohio State University, Columbus, Ohio, 2008.

Terekhov, A.S., 1975, Basic Problems of Heat Calculation of Gear Reducers, JSME
International Conference on Motion and Powertransmissions, Nov. 23-26, 1991, 490495.

van Dongen, L. A. M., 1982, Efficiency Characteristics of Manual and Automatic


Passenger Car Transaxles, SAE Technical Paper Series, 820741, Warrendale, PA:
Society of Automotive Engineers.

Von Karman, T., 1921, On Laminar and Turbulent Friction, Z. Angew. Math. Mech., 1,
235-236.

Wild, P. M., Dijlali, N., and Vickers, G. W., 1996, Experimental and Computational
Assessment of Windage Losses in Rotating Machinery, ASME Trans. J. Fluids Eng.,
118, 116122.

Xu, H., Kahraman, A., Anderson, N.E., Maddock, D.G., 2007, Prediction of Mechanical
Efficiency of Parallel-Axis Gear pairs, ASME Journal of Mechanical Design, 129, 5868.

169

You might also like