DYNAMIC MODELING OF DOUBLE-HELICAL PLANETARY SETS

DISSERTATION DRAFT

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

By

Sondkar Prashant, M.Tech.

Graduate Program in Mechanical

The Ohio State University 2012

Dissertation Committee:

Dr. Ahmet Kahraman, Advisor

Professor Daniel Mendelsohn

Professor Manoj Srinivasan

Professor Junmin Wang

© Copyright by

Sondkar Prashant

2012

ABSTRACT

This dissertation aims at investigating the dynamic response of double-helical

planetary gear sets theoretically. A three-dimensional discrete dynamic model of a

double-helical planetary gear set is proposed, including all gear mesh, bearing and

support structure compliances. The model is presented in three levels of complexity: (i) a

linear time-invariant (LTI) model, (ii) a LTI model with gyroscopic effects included, and

(iii) a nonlinear time-varying (NTV) model with parametrically time-varying gear mesh

stiffnesses and nonlinear tooth separation effects included.

As the first step, a generic linear (no tooth separations), time-invariant (constant

gear mesh stiffnesses) dynamic model is formulated to analyze any N-planet double-

helical planetary gear system. The model includes any planet phasing conditions dictated

by the number of planets, number of gear teeth and planet position angles as well as any

phase shifts due to the designed stagger between the right and left sides of the gear set.

The forced response due to gear mesh errors excitations is computed by

using the modal summation technique with the natural modes found from the

corresponding Eigen value problem for the undamped system. The strain energies of the

mode shapes are computed to identify the modes excitable by these excitations.

Parametric studies are presented to demonstrate sizable influences of planet phasing,

ii

stagger conditions, gear and carrier support conditions as well as the number of planets

on the steady-state forced response.

In the second modeling step, the LTI model is modified to include a class of

gyroscopic effects due to vibratory skew of spinning for the case of a stationary

carrier. The complex Eigen solutions are examined to quantify the influence of rotational

speed of the gear set through gyroscopic effects on the natural modes. A complex modal

summation formulation is used to compute the forced response with gyroscopic effects.

Results indicate that the influence of gyroscopic moments on natural frequencies is

modest within typical speed ranges, with only a sub-set of modes exhibiting dominant

tilting motions impacted by the gyroscopic effects. Effect of gyroscopic moments on

forced response curves is found to be limited to slight changes in amplitudes and

frequencies of certain resonance peaks.

As the final step, mesh stiffness fluctuations due to change in number of tooth pairs

are introduced as internal parametric excitation along with the transmission error

excitations at the same phasing relations. Tooth separation functions are also applied to

obtain a set of NTV equations of motion, which are solved by using direct numerical

integration. Differences observed between the forced response curves for time-varying

and time-invariant systems are characterized by additional resonance peaks and overall

increases in response amplitudes while no signs of nonlinear behavior are noted.

iii

DEDICATION

Dedicated to

My dear family

iv

ACKNOWLEDGMENTS

I would like to express my sincere gratitude to Prof. Ahmet Kahraman for the opportunity, guidance, and support throughout my research at The Ohio State University.

I am grateful to Prof. Daniel Mendensohn, Prof. Junmin Wang, and Prof. Manoj

Srinivasan for serving on my dissertation committee.

Special thanks to Mr. Jonny Harianto for providing software support. I am thankful to Mr. Sam Shon and Dr. David Talbot for their technical expertise and willingness to share it. I would also like to extend my thanks to all my lab mates, including, but not limited to Devin Hilty and Mohammad Hotait for their friendship and support throughout my work at Gear Lab. I am thankful to Pratt & Whitney for sponsoring this research activity.

I want to sincerely thank my dear family for their continuous support and encouragement without which this work would not have been possible.

v

VITA

June 2004 ...... Bachelor of Engineering Pune University, India

June 2006 ...... Master of Indian Institute of Technology (IIT), Madras, India

2005-2006 ...... DAAD Fellowship Technical University of Darmstadt, Germany

2006-2008 ...... Engineer (Engine Dynamics) General Electric, India

2008-2012 ...... Graduate Research Associate The Ohio State University, OH

FIELDS OF STUDY

Major Field: Mechanical Engineering

Focus on Gear Dynamics.

vi

TABLE OF CONTENTS

ABSTRACT ...... ii

DEDICATION ...... iv

ACKNOWLEDGMENTS ...... v

VITA ...... vi

LIST OF TABLES ...... xi

LIST OF FIGURES ...... xii

NOMENCLATURE ...... xv

CHAPTERS:

1. Introduction ...... 1

1.1 Background and Motivation ...... 1

1.2 Literature Review ...... 6

1.3 Scope and Objectives ...... 11

1.4 Dissertation Outline ...... 13

References for Chapter 1 ...... 15

vii

2. A Linear Time-invariant Dynamic Model of a Double-Helical Planetary Gear Set ...... 20 2.1 Introduction ...... 20

2.2 Discrete Model and its Assumptions ...... 21

2.2.1 A Sun-Planet i Pair Formulation ...... 23

2.2.2 A Ring-Planet i Pair Formulation ...... 29

2.2.3 A Carrier-Planet i Pair Formulation ...... 34

2.2.4 of the Left and Right Sides ...... 38

2.2.5 The Overall System Equations ...... 41

2.2.6 Excitations ...... 43

2.3 Solution Methodology ...... 48

2.4 An Example Simulation ...... 51

2.4.1 Influence of Right-to-left Stagger ...... 54

2.4.2 Influence of Planet Phasing Conditions ...... 65

2.4.3 Influence of Number of Planets ...... 68

2.4.4 Influence of Radially Floating Sun Gear ...... 71

2.5 Mode Identification using Modal Strain Energy ...... 73

2.6 Summary ...... 80

References for Chapter 2 ...... 81

3. Influence of Gyroscopic Effects on Dynamic Behavior of Double-Helical Planetary Gear Sets ...... 83 3.1 Introduction ...... 83

3.2 Incorporation of Gyroscopic Moments in the Dynamic Model ...... 84

3.2.1 A Sun-Planet i Pair with Gyroscopic Effects ...... 85

viii

3.2.2 A Ring-Planet i Pair with Gyroscopic Effects ...... 87

3.2.3 A Carrier-Planet i Pair with Gyroscopic Effects ...... 88

3.2.4 Coupling Elements with Gyroscopic Effects ...... 89

3.2.5 The Overall System Equations with Gyroscopic Effects ...... 90

3.3 Solution Methodology ...... 91

3.4 An Example Simulation ...... 94

3.5 Summary ...... 103

References for Chapter 3 ...... 105

4. Investigation of Time-Varying Gear Mesh Stiffness Effects on Dynamics of Double-Helical Planetary Gear Sets ...... 106 4.1 Introduction ...... 106

4.2 A Nonlinear Time-Varying Dynamic Model ...... 109

4.2.1 Definition of Time-Varying Gear Mesh Stiffnesses ...... 109

4.2.1 Equations of Motions for NTV Model ...... 112

4.3 Solution of the NTV System Equations ...... 115

4.4 Numerical Results ...... 117

4.4.1 Verification and Analysis of Time-domain Solutions ...... 118

4.4.2 Example System Analyses ...... 119

4.5 Summary ...... 131

References for Chapter 4 ...... 132

5. Conclusions and Future Recommendations ...... 135

5.1 Summary ...... 135

5.2 Conclusions and Contributions ...... 137 ix

5.3 Recommendations for Future Work ...... 140

BIBLIOGRAPHY ...... 141

Appendix A Beam Element Matrices ...... 148

Appendix B Overall System Matrices ...... 153

Appendix C Elements of Forcing Vector ...... 158

x

LIST OF TABLES

Table Page

2.1 Basic design parameters of the example gear system ...... 52

2.2 Harmonic amplitudes and phase angles of the transmission error excitations of the example gear set of Table 2.1 [2.6] ...... 53

2.3 Predicted natural frequencies and mode types of the example gear set ...... 55

2.4 Strain Energy distribution for the modes of the example gear set (excited modes are shown in italic characters) ...... 78

3.1 Strain energy distribution for the modes exhibiting change in natural frequencies due to gyroscopic effects ...... 96

4.1 Harmonic amplitudes and phase angles of the transmission error and mesh stiffness excitations of the example gear set of Table 2.1 [4.19] for different torque levels ...... 121

xi

LIST OF FIGURES

Figure Page

1.1 Main components of planetary gear set ...... 2

1.2 An example double helical planetary gear set [1.2] ...... 4

2.1 Dynamic model of a double-helical planetary gear system ...... 24

2.2 Dynamic model of sun-planet i pair ...... 25

2.3 Dynamic model of ring-planet i pair ...... 30

2.4 Dynamic model of carrier-planet i pair ...... 35

2.5 (a) Geometry of a double helical external gear, and (b) three-piece model of the double helical gear ...... 40

2.6 Illustration of the right-to-left stagger conditions in a double helical gear pair; (a) 0 and (b)  ...... 46 stg stg 2.7 Maximum dynamic mesh force amplitudes at the left side (a) s-pi and (b) r-pi meshes ...... 56

2.8 Maximum dynamic mesh force amplitudes at the right side (a) s-pi and (b) r-pi meshes ...... 57

2.9 Mode shapes representative of Eq. (2.32) at (a) 2 1262 Hz and (b) 38 11047 Hz ...... 60

2.10 Mode shape representative of Eq. (2.33) at (a) 14 2994 Hz and (b) 15  3166 Hz ...... 61

2.11 Dynamic factors at the left side (a) s-pi and (b) r-pi meshes ...... 63

2.12 Dynamic factors at the right side (a) s-pi and (b) r-pi meshes ...... 64

xii

2.13 Maximum dynamic mesh force amplitudes at the left/right side s-pi meshes for (a) 0 and (b)  for different stg stg planet phasing conditions ...... 69

2.14 Maximum dynamic mesh force amplitudes at the left/right side (a) s-pi and (b) r-pi meshes for different number of planet gears ...... 70

2.15 Effect of radially floating sun gear on dynamic mesh force amplitudes for (a) stg 0 and (b) stg ...... 72

2.16 Normalized modal strain energy components of modes at (a)  8398 Hz (b)  9451Hz ...... 79

3.1 Variation of certain natural frequencies with the rotational speed due to gyroscopic effects (C0 ) ...... 95

3.2 Strain energy distribution for support spring and planet bearings for modes at (a) 7 2076 Hz and (b) 14  2994 Hz. (Tr - Translational, Ti - Tilting, Ax - Axial) ...... 98

3.3 Maximum percent change in natural frequencies with an input speed range of 0 to 20,000 rpm as a function of (a) sun gear support stiffness in tilting direction and (b) ring gear support stiffness in tilting direction ...... 99

3.4 Maximum percent change in natural frequencies with an input speed range of 0 to 20,000 rpm as a function of polar mass moment of inertia of (a) the sun gear and (b) the ring gear ...... 100

3.5 Maximum dynamic mesh force amplitudes at the left side (a) s-pi and (b) r-pi meshes with and without gyroscopic effect ...... 102

3.6 Maximum dynamic mesh force amplitudes at the left side (a) s-pi and (b) r-pi meshes with and without gyroscopic effect for radially floating sun ( stg ) ...... 104

4.1 Comparison of modal summation and direct numerical integration solutions for the LTI case. Maximum dynamic mesh force amplitudes on the left side (a) s-pi and (b) r-pi meshes ...... 120

4.2 Root-mean-square. values of dynamic mesh forces at the left side (a) s-pi and (b) r-pi meshes for different excitation models ( stg 0 ) ...... 123

xiii

4.3 Steady-state dynamic mesh force at an s-pi mesh on the left side obtained by using Model II. (a) Time history at m  8852 Hz and (b) the corresponding frequency spectrum ...... 124

4.4 Steady-state dynamic mesh force at an s-pi mesh on the left side obtained by using Model II. (a) Time history at m  5523Hz and (b) the corresponding frequency spectrum ...... 125

4.5 Maximum dynamic mesh force amplitudes on the left side of (a) s-pi and (b) r-pi meshes with different excitation models ( stg 0 ) ...... 126

4.6 Maximum dynamic mesh force amplitudes at the left side (a) s-pi and (b) r-pi meshes for different excitation models ( stg ) ...... 128

4.7 Maximum dynamic mesh force amplitudes on the left side (a) s-pi and (b) r-pi meshes of an in-phase system for different gear mesh models ( stg 0 ) ...... 129

4.8 Maximum dynamic mesh force amplitudes on the left side of (a) s-pi and (b) r-pi mesh at various input torque levels. Model II was used here ...... 130

A.1 An Euler beam element ...... 152

xiv

NOMENCLATURE

Symbol Definition

C Damping matrix

D Diameter e Transmission error eˆ Amplitude of l-th harmonics of transmission error

F Force acting on planet i due to coupling between carrier and planet i

Fspi Dynamic mesh force for sun-planet i gear mesh

Frpi Dynamic mesh force for ring-planet i gear mesh

F Force vector

FW Face width

FWg Length of gap between two sides of double helical gears

G Gyroscopic matrix h Unite step function representing tooth separation

H Angular momentum vector

I Diametral mass moment of inertia

J Polar mass moment of inertia

xv k Planet bearing stiffness ksp Mean value of sun-planet i gear mesh stiffness krp Mean value of ring-planet i gear mesh stiffness kx Support bearing stiffness in x direction (   src,, ) ky Support bearing stiffness in y direction (   src,, ) kz Support bearing stiffness in z direction (   src,, ) kx Support bearing stiffness in x direction (   src,, ) ky Support bearing stiffness in y direction (   src,, ) kˆ Amplitude of l-th harmonics of mesh stiffness

KD Dynamic factor

L Modal matrix of left eigen vector m Mass

M Moment acting on planet i due to coupling between carrier and planet i

M Mass matrix

N Number of Planets p Relative mesh displacement pˆ Modal relative mesh displacement q Forced vibration response vector

Q Mode shape r Base radii of gear  (   sr,, pi) xvi rc Planet pin hole radius r State vector

R Modal matrix of right eigen vector

T Torque

Tmesh Mesh period

U Strain energy x Forced vibration response in x direction xˆ Modal displacement in x direction y Forced vibration response in y direction yˆ Modal displacement in y direction z Forced vibration response in z direction zˆ Modal displacement in z direction

Z Number of teeth on gear

 pi Position angle for planet i

 Helix angle

spi phase angle between the s-pi mesh on the left side and the reference s-p1 mesh

rpi phase difference between the r-pi mesh and the r-p1 mesh on the left side

stg phase difference between the left and right side due to intentional nominal stagger of the teeth

rs phase difference between the reference s-p1 mesh and the r-p1 mesh on left side

xvii

 Stiffness multiplier for proportional damping

 Damping factor

 Mass multiplier for proportional damping

x Forced vibration response in x direction

y Forced vibration response in y direction

z Forced vibration response in z direction

ˆ x Modal displacement in x direction

ˆ x Modal displacement in y direction

ˆ z Modal displacement in z direction

 Dynamic compliance

 Phase angle of l-th harmonics of transmission error excitation

 Phase difference

 Phase angle of l-th harmonics of mesh stiffness

 Transvers pressure angle

 Angle made by plane of action with vertical y axis

 Natural frequency

m Mesh frequency

 Velocity vector

 Rotational speed

xviii

Subscripts b Bore c Carrier i Index for number of planets l Harmonics number m Mean value mesh Gear mesh n Number of beam element r Ring gear pb Planet bearing pi Planet i gear s Sun gear sup Support bearing x x direction y y direction z z direction

x x direction

y y direction

z z direction

 Number of degrees of freedom

xix

Superscript

L Left side of double helical gear

R Right side of double helical gear

xx

CHAPTER 1

Introduction

1.1 Background and Motivation

Planetary gear sets (also known as epicyclic gear sets) are commonly used in many automotive, industrial, aerospace and industries. The typical applications include jet propulsion systems, rotorcraft transmissions, passenger , wind turbine gearboxes and other industrial gearboxes. Figure 1.1 shows the components of a simple planetary gear set consisting of an N number of planet gears

(typically N [3,7] ) that are in mesh with an external sun gear (s) and an internal ring

gear (r). The planet gears are supported by a rigid structure called carrier (c) through various types of planet bearings (needle bearings, rolling element bearings, and in some aerospace applications, journal bearings).

One of the main advantages of planetary gear sets is that the input power is split

into a number of parallel power transmission paths, providing higher power density

(power to weight ratio) values. This lowers the forces carried by individual planet meshes to allow smaller tooth modules with higher gear contact ratios, resulting in quieter gear 1

carrier carrier planet planet

planet pin

rolling element planet pin bearing ring

Figure 1.1: Main components of planetary gear set

2

set designs [1.1]. Different speed ratios can be obtained from the same gear set by

changing input, output and stationary (reaction) members. This capability is the basis for

their extensive use in automotive automatic transmissions and transfer cases. In addition,

the axi-symmetric (coaxial) configuration of planetary gear sets eliminates the most of

the radial loads on bearings [1.1]. Furthermore, planetary gear sets have the ability to

self-center themselves radially to compensate for various manufacturing errors such as

carrier pin-hole position errors and gear eccentricities.

Planetary gear sets consist of either spur, (single) helical or double-helical gears.

Spur planetary gear sets can be commonly found in heavy machinery and off-highway

gearboxes and transmissions, while the helical planetary gear sets are the norm for all

automotive applications as in automatic transmissions and transfer cases. When helical

gears are used in a planetary gear set, axial thrust forces are created on the sun and ring

gears that must be balanced by bearings or the thrust forces on the adjacent planetary gear sets. Double-helical planetary gear sets, as the one shown in Figure 1.2 [1.2] for a jet engine turbofan application, do not have this problem as the static axial gear mesh forces acting on the right and left sides cancel each other. In addition, they have higher load carrying capacity and better noise characteristics. For these reasons, they have been used for jet engine and rotorcraft gearbox applications in spite the cost and manufacturing challenges associated with their production.

Planetary gear sets have several unique issues/features associated with their functional behavior. As each planet branch is a split power transmission path, an equal planet-to-planet load sharing is often not possible especially when certain manufacturing 3

Figure 1.2: An example double helical planetary gear set [1.2].

4

tolerances of the gear and the carrier are not designed to be very tight and all of the central members (sun, ring or carrier) are supported rigidly in the radial direction [1.3, 1.

4]. Reduced gear thicknesses not only reduce the mass but also shown to reduce the influence of internal gear and carrier errors [1.5]. It is also reported that a flexible internal gear improves the equal load sharing amongst the planets. The structural modal properties of planetary gears with equally spaced planets possess choices of certain phasing conditions such that cancellation or neutralization of dynamic gear mesh forces can be achieved through design strategies to reduce vibration and noise levels. Proper planet mesh phasing can eliminate many troublesome vibration modes [1.6]. Planetary gear sets also exhibit unique vibration frequency spectra with amplitude or frequency modulated sideband components associated either with the carrier rotation or with the eccentricities of gear components [1.7-1.9]. These vibration signatures were exploited for fault diagnosis purposes like detection of cracks in the carrier [1.10].

Dynamic analysis of a double-helical planetary gear set, as any other gear system, is required for two main reasons. One reason is associated with the resultant noise outcome. High-frequency dynamic forces created at the gear meshes are transmitted to gearbox housing and the surrounding support structures to generate structure-borne noise.

Therefore, reduction of vibrations necessitates a better understanding of the dynamic behavior. The other main reason stems from the durability and reliability requirements of the planetary gear set. Dynamic gear mesh and bearing forces that alternate about the static forces transmitted cause dynamic stresses and dynamic factors to impact the fatigue lives of the gears and bearings adversely. Hence, dynamic analysis of a double helical

5

planetary gear set becomes a critical step in the design process, which aims at developing

quiet and reliable transmissions.

In this research, a theoretical will be performed to understand the dynamic behavior

of double helical planetary gear sets. Towards this aim, an analytical modeling

framework will be formulated and executed, which will serve as a fundamental tool for

design and further research. The proposed models will be used to investigate the free and

forced vibration response of the double helical planetary gear sets. Gyroscopic effects,

time varying mesh stiffness and gear mesh excitations will all be incorporated in the

model.

1.2 Literature Survey

Numerous analytical models on planetary gear dynamics consisting of spur or

helical gears can be found in the literature while research on double-helical gears is quiet

limited. In their review paper, Yang and Dai [1.11] presented an overview of various

planetary gear dynamics models found in literature, most of which are focused on

analytical work comprising of discrete-parameter models with gears bodies assumed to

be rigid and tooth flexibilities are modeled as a springs. These models vary in terms of

degrees of freedom included (purely torsional, two-dimensional, or three-dimensional)

and type of formulation employed (linear time-invariant (LTI) models with constant gear

mesh stiffnesses and no backlash, linear time-varying models (LTV) with fluctuating

mesh stiffnesses and no gear backlash, or nonlinear time-varying models (NTV) with

both fluctuating mesh stiffnesses and backlash included). 6

Types of analyses performed in these studies also vary as some focused on

prediction of natural frequencies and mode shapes, others emphasized the excitation

and cancellation/minimization of these excitations by properly phasing the

planets while a number of studies predicted forced vibration response and dynamic tooth

load. Botman [1.12] carried out a study to investigate the modes of planetary gear set

with a non-rotating or rotating carrier. Kahraman [1.13] developed a torsional model of

planetary gear set and established closed-form expressions for natural frequencies and

mode shapes of planetary system consisting of any number of planets. A dynamic model

for planetary gear set was presented by Saada and Velex [1.14] to study the influence of

gear mesh stiffnesses and support stiffnesses on natural frequencies of the system.

Kahraman [1.1] was first to investigate the free torsional vibration characteristics of a

compound planetary gear set. Effort was made to classify modes as rigid body mode,

asymmetric planet modes and axi-symmetric overall modes. Inalpolat and Kahraman

[1.15] developed a model to study free and forced vibration characteristics of any

planetary formed by N stages of different types (single-planet, double-planet, or

complex-compound).

Studies on reducing the dynamic mesh forces by proper planet phasing to cancel or

neutralize the mesh excitations goes back to Seager [1.6] who established conditions for

neutralization of harmonic components of excitations of the central members of a

planetary gear set formed by spur gears, which was achieved by suitable choice of number of gear teeth. Toda and Botman [1.16] showed that the vibration excitations can be reduced significantly with relative indexing of planets. They focused their effort on

7

establishing how indexing of planets can minimize the effect of tooth spacing errors in

the planets. Kahraman and Blankenship [1.17] provided the first generalized phasing

formulation, defining the relationships amongst all of the gear meshes of an N-planet,

helical planetary gear set. This formulation showed that there is no specific phasing

condition that can neutralize the excitations in all directions, but it is possible to find a

phasing condition that can yield a desirable response for certain applications, operating

within a given speed range. Parker and Lin [1.18] provided the analytical description of

mesh phasing relationships between sun-planet and ring-planet meshes. Platt and

Leopold [1.19] carried out experiments to study the effect of planet mesh phasing on

noise levels. Different phasing conditions including in-phase and sequentially phased

planets were investigated for their respective noise levels.

Third group of studies focused on prediction of the forced vibration response and

dynamic gear mesh and bearing forces. Cunliffe et al [1.20] studied the variation of

dynamic tooth load with planet pin stiffness for single stage planetary gear set consisting

of spur gears. Hidaka and Terauchi [1.21] conducted experiments to study the dynamic

load sharing among planet meshes and compared the results with analytical model [1.22].

Influence of run-out errors on the motion of floating sun gear was investigated by Hidaka

et al [1.23]. August and Kasuba [1.24] developed a torsional model of single stage

planetary gear system with shafts, input and output units. Effect of a floating, partially

floating and fixed sun gear on dynamic mesh forces was investigated by incorporating

additional transverse degrees of freedom for the sun gear. Motion of sun gear was found

to be influenced by mesh stiffness variation as well. Kahraman [1.25] presented a two-

8

dimensional NTV model of a planetary gear set. This study is focused on investigating

the influence of design, manufacturing and assembly variations on dynamic planet-to-

planet load sharing characteristics.

The first generalized three-dimensional model for a helical planetary gear set was

proposed by Kahraman [1.26]. This LTI model included the planet phasing formulation

of Kahraman and Blankenship [1.17] as well as all six degrees of freedom for each gear

and the carrier. A purpose of this model was to simulate the dynamic loads at individual

gear meshes in relation to particular meshing conditions to identify the best possible

phasing configurations. Velex and Flamand [1.27] extended their earlier modeling effort

[1.14] to include the contact formulation in the dynamic model of a planetary gear

system. A numerical method was introduced to define the instantaneous positions of

contact and corresponding mesh parametric excitations. Time integration method was

employed to solve the system of equations. A generalized torsional NTV model was

presented by Al-Shyyab and Kahraman [1.28] by including gear backlash type clearance

nonlinearities and time-varying stifffnesses. The multi-term Harmonic Balance Method

was used to solve nonlinear equations of motion analytically. Chaari et al [1.29]

investigated the effect of manufacturing errors on the dynamic behavior of planetary gears. Influence of eccentricity of gears and profile errors on the frequency response of system was studied. Parametric instability of planetary gear system due to variable mesh stiffness was analyzed by Hbaieb et al [1.30]. A rectangular waveform was used to represent the mesh stiffness variations. A perturbation technique was employed to solve

9

the system of equations. Stability boundaries were identified, which included primary,

secondary and combination instabilities.

In recent years, some researchers have used deformable gear body dynamic models.

Parker et al [1.31] employed contact model developed by Vijayakar [1.32] to build deformable body dynamic model of planetary gear system and studied the dynamic behavior of the system over a range of speeds. Effects of gear rim thickness parameters

on the gear stresses were analyzed by Kahraman et al [1.5]. Yuksel and Kahraman [1.33]

combined a surface wear model and deformable-body dynamic model to study the

influence of tooth wear on dynamic mesh forces. A hybrid three dimensional lumped

parameter and finite element model was developed by Abousleiman et al [1.34] to

investigate the effect of flexible ring gear on the dynamic response. The same

investigators published another study [1.35] to look at the effects of geometrical errors

and centrifugal forces on the dynamic response.

Literature on double helical gears is quiet sparse and limited to a single gear pair

arrangement, not multi-mesh systems such as planetary gear sets. Thomas [1.36]

developed an analytical model for investigating load distribution and transmission error

of a double helical gear pair under quasi-static conditions. The model was later used by

Clapper and Houser [1.37] to investigate the root stresses of double helical gears and

comparison was made with experiments performed. Zhang et al [1.38] carried out a noise

optimization of double helical parallel shaft gearbox by developing a three-dimensional

FE model. Noise reduction was achieved by varying the thickness of internal bearing

supporting panels and external walls of gearbox. Wang et al [1.39] presented study about 10

tooth modification of double helical gear pair to reduce the transmission error.

Experiments were performed on optimized tooth geometry to study its transmission error characteristics. Jauregui and Gonzalez [1.40] developed single degree-of-freedom model to study axial vibrations of double helical gear pair due to manufacturing errors. Quasi- static and dynamic analyses of a double helical gear pair was carried out by Ajmi and

Velex [1.41]. A 12-DOF model of a left side of gear pair was combined with another 12-

DOF model of the companion right side pair using Euler beam elements. The effect of floating and staggering of teeth on the quasi-static and dynamic behavior of a gear pair was investigated. Anderson et al [1.42] conducted experiments on double helical planetary gear set to measure efficiency, vibration amplitudes and stress levels.

1.3 Scope and Objectives

The literature review presented above reveals that the dynamic behavior of double- helical planetary gear sets has not been studied. Modeling of double-helical planetary gear sets involve additional complications over (single) helical planetary gear sets in terms of left-to-right side load sharing and left-to-right gear mesh phasing relations. It is evident that a methodology to analyze double helical gears does not exist and no published model can be found in the literature describing dynamic characteristics of double-helical planetary gear systems. In fact, double helical systems attracted very little attention even in terms of their single gear pair dynamic behavior. Accordingly, overall objective of this dissertation is to develop analytical models to study essential dynamic characteristics of a double helical planetary gear system.

11

The models will be based on a discrete representation of the planetary gear set. All

gear bodies will be assumed to be rigid disks with gear tooth compliance represented by

spring and damper elements. The same discrete treatment will be applied to bearings as

well. As the damping mechanisms of such systems are the least known, modal and proportional damping formulations will be used. The dissertation objective stated above

will be achieved in three steps of increasing complexity:

(i) Development of a Linear Time Invariant (LTI) model: A new LTI model of a

double helical planetary gear set will be developed to predict the free vibration

characteristics and forced vibration response under simplified damping

conditions. This generalized three-dimensional model will include any

number of equally or unequally positioned planets, as well as any planet

phasing and support conditions. Torsional, transverse, axial and rotational

(tilting) motions of gears and the planet carrier will be included in this three-

dimensional model. A detailed parametric study will be performed on this

model to investigate the influence of different parameters on the dynamic

response.

(ii) Incorporation of Gyroscopic Effects: The LTI model will be expanded to

include high speed effects such as gyroscopic moments to study their

influence on the dynamic behavior of the system. Sensitivity of Eigen values

and the forced response to rotational speed and gyroscopic effects will be

investigated.

12

(iii) Development of a Nonlinear Time Varying (NTV) Model: In order to quantify

the nonlinear and time varying effects, periodically varying mesh stiffnesses

due to variable number of teeth in contact and tooth separation nonlinearity

will be incorporated in the earlier LTI model. Furthermore, a piecewise

clearance function will be incorporated, to model any tooth separations.

Numerical time integration scheme will be adopted to solve the system of

nonlinear equations.

The purpose of this research is to gain insight into the dynamic behavior of double helical planetary system with simplified model without resorting to deformable-body modeling of gear sets, due to computational limitations. Support structures and bearings will modeled in a simple manner as done in most of previous gear dynamic models.

1.4 Dissertation Outline

As stated in the previous section, the modeling effort will be carried out in three steps. Each of these steps will be described in an individual chapter. Chapter 2 is focused on development of the LTI model of a double helical planetary gear set. The modeling assumptions along with details of the modeling methodology will be presented.

The equations of motion will be derived and solved to predict natural modes and the forced response of the gear set within a given range of operating speed. This model will serve as the foundation for the other models. A parametric study will be carried out to investigate the influence of different design parameters and variations such as right-to-

13

left gear tooth stagger, support stiffnesses, planet position angles, planet mesh phasing

conditions and number of planets on the dynamic behavior of the system.

In Chapter 3, the LTI formulation of Chapter 2 will be modified to include gyroscopic moments. Gyroscopic effects are often hypothesized to be important in high- speed gear applications in aerospace industry with little work to substantiate it.

Inclusion of gyroscopic effects will result in an asymmetric damping matrix. The governing complex Eigen Value problem will be solved to quantify the combined effect of speed and gyroscopic moments on natural modes of the planetary gear set. The modal summation technique will be used to determine the forced response. Direct comparisons between the cases when the gyroscopic effect included and ignored will be presented to determine conditions when gyroscopic models must be included in the model.

Chapter 4 provides a further expansion of earlier models by including periodic time variation of gear mesh stiffnesses as well as contact loss induced by the backlash present at the gear meshes. As inclusion of these effects makes the stiffness matrix a periodically time-varying one, subject to a nonlinear backlash (clearance) constraint, the frequency- domain (modal summation) solutions of the previous chapters are not applicable to this

NTV system. Gyroscopic effects will also be included in this NTV model. The system equations will be put into the state-space form and solved by the direct numerical integration method. The results of the NTV model will be compared to the corresponding LTI models with and without gyroscopic effects to determine whether time-varying gear mesh stiffness effects and gear backlash should be included in modeling of double helical planetary gear sets. 14

Finally, Chapter 5 summarizes entire work, and lists major conclusions and the

contributions of the proposed research. Recommendations for future research on this

topic to improve the modeling effort are also included in this chapter.

References for Chapter 1:

[1.1] Kahraman, A., 2001, “Free Torsional Vibration Characteristics of Compound Planetary Gear Sets,” Mechanism and Theory, 36, pp. 953-971.

[1.2] Mraz, S., 2009, “Gearing up for Geared Turbofan,” Machine Design by Engineers for Engineers. (http://machinedesign.com/article/gearing-up-for-geared-turbofans- 0202)

[1.3] Ligata, H. and Kahraman, A., 2008, “An Experimental Study of the Influence of Manufacturing Errors on the Planetary Gear Stresses and Planet Load Sharing,” ASME Journal of Mechanical Design, 130, 041701-1 – 041701-9.

[1.4] Bodas, A. and Kahraman, A., 2004, “Influence of Carrier and Errors on the Static Load Sharing behavior of Planetary Gear Sets,” JSME International Journal, 47, pp. 908-915.

[1.5] Kahraman, A., Kharazi, A., and Umrani, M., 2003, “A Deformable Body Dynamic Analysis of Planetary Gears with Thin Rims,” Journal of Sound and Vibration, 262, pp. 752-768.

[1.6] Seager, D., 1975, “Conditions for the Neutralization of Excitation by the Teeth in Epicyclic Gearing,” Journal Mechanical Engineering Science, 17(5), pp. 293- 298.

[1.7] McFadden, P. and Smith, J., 1985, “An Explanation for the Asymmetry of the Modulation Sidebands about the Tooth Meshing Frequency in Epicyclic Gear Vibration,” Proceedings of Institution of Mechanical Engineers, 199(C1), pp. 65- 70.

15

[1.8] Inalpolat, M. and Kahraman, A., 2009, “A Theoretical and Experimental Investigation of Modulation Sidebands of Planetary Gear Sets,” Journal of Sound and Vibration, 323, pp. 677-696.

[1.9] Inalpolat, M. and Kahraman, A., 2010, “A Dynamic Model to predict Modulation Sidebands of a Planetary Gear Set having Manufacturing Errors,” Journal of Sound and Vibration, 329, pp. 371-393.

[1.10] Blunt, D. and Keller, J., 2006, “Detection of Fatigue Crack in a UH-60A Planet Gear Carrier using Vibration Analysis,” Mechanical Systems and Signal Processing, 20, pp. 2095-2111.

[1.11] Yang, J. and Dai, L., 2008, “Survey of Dynamics of Planetary Gear Trains,” International Journal of Materials and Structural Integrity, 1, pp. 302-322.

[1.12] Botman, M. 1976, “Epicyclic Gear Vibrations,” Journal of Engineering for Industry, 97, pp. 811-815.

[1.13] Kahraman, A., 1994, “Natural Modes of Planetary Gear Trains,” Journal of Sound and Vibration, 173(1), pp. 125-130.

[1.14] Saada, A. and Velex, P., 1995, “An Extended Model for the Analysis of the Dynamic Behavior of Planetary Trains, ” ASME Journal of Mechanical Design, 117, pp. 241-247.

[1.15] Inalpolat, M. and Kahraman, A., 2008, “Dynamic Modeling of Planetary Gears of Automatic Transmissions,” Proceedings of Institution of Mechanical Engineers Part K: Journal of Multi-body Dynamics, 222, pp. 229-242.

[1.16] Toda, A. and Botman, M., 1979, “Planet Indexing in Planetary Gears for Minimum Vibrations,” ASME paper, 79-DET-73.

[1.17] Kahraman, A. and Blankenship, G., 1994, “Planet Mesh Phasing in Epicyclic Gear Sets,” Proceedings of ASME Power Transmission and Gearing Conference, San Diego.

16

[1.18] Parker, R. and Lin, J., 2004, “Mesh Phasing Relationships in Planetary and Epicyclic Gears,” ASME Journal of Mechanical Design, 126, pp. 365-370.

[1.19] Platt, R. and Leopold, R., 1996, “A Study on Helical Gear Planetary Phasing Effects on Transmission Noise,” VDI Berichte, 1230, pp.793-807.

[1.20] Cunliffe, F., Smith, J., and Welbourn, D., 1974, “Dynamic Tooth Loads in Epicyclic Gears,” ASME Journal of Engineering for Industry, 95, pp. 578-584.

[1.21] Hidaka, T. and Terauchi, Y., 1976, “Dynamic Behavior Planetary Gear, 1st Report: Load Distribution in Planetary Gear,” Bulletin of JSME, 19, pp. 690-698.

[1.22] Hidaka, T., Terauchi, Y., and Fujii, M., 1980, “Analysis of Dynamic Tooth Load on Planetary Gear,” Bulletin of JSME, 23, pp. 315-323.

[1.23] Hidaka, T., Terauchi, Y., and Dohi, K., 1979, “On the Relation between the Run- Out Errors and the Motion of the Center of Sun Gear in Stoeckicht,” Bulletin of JSME, 22(167), pp. 748-754.

[1.24] August, R. and Kasuba, R., 1986, “Torsional Vibration and Dynamic Loads in a Basic Planetary Gear System,” ASME Journal of Vibration, Acoustics, Stress, and Reliability in Design, 108, pp. 348-353.

[1.25] Kahraman, A., 1994, “Load Sharing Characteristics of Planetary Transmission,” Mechanism and Machine Theory, 29(8), pp. 1151-1165.

[1.26] Kahraman, A., 1994, “Planetary Gear Train Dynamics,” ASME Journal of Mechanical Design, 116, pp. 713-720.

[1.27] Velex, P. and Flamand, L., 1996, “Dynamic Response of Planetary Trains to Mesh Parametric Excitations,” ASME Journal of Mechanical Design, 118, pp. 7- 14.

[1.28] Al-Shyyab, A. and Kahraman, A., 2007, “A Nonlinear Dynamic Model for Planetary Gear Sets,” Proceedings of the Institution of Mechanical Engineers Part K: Journal of Multi-Body Dynamics, 221, pp. 567-576.

17

[1.29] Chaari, F., Fakhfakh, T., Hbaieb, R., Louati, J., and Hadder, M., 2006, “Influence of Manufacturing Errors on the Dynamic Behavior of Planetary Gears,” International Journal of Advance manufacturing Technology, 27, pp. 738-746.

[1.30] Hbaieb, R., Chaari, F., Fakhfakh, T., Hadder, M., 2006, “Dynamic Stability of a Planetary Gear Train under the Influence of Variable Meshing Stiffnesses,” Proceedings of IMechE. Part D: Journal of Automobile Engineering, 229(D12), I711-I725.

[1.31] Parker, R., Agashe, V., and Vijayakar, S., 2000, “Dynamic Response of a Planetary Gear System using a Finite Element/Contact Mechanics Model,” ASME Journal of Mechanical Design, 122, pp. 304-310.

[1.32] Vijayakar, S., 1991, “A Combined Surface Integral and Finite Element Solution for a Three-Dimensional Contact Problem,” International Journal for Numerical Methods in Engineering, 31, pp. 525-545.

[1.33] Yuksel, C. and Kahraman, A., 2004, “Dynamic Tooth Loads of Planetary Gear Sets having Tooth Profile Wear,” Mechanism and Machine Theory, 39, pp. 695- 715.

[1.34] Abousleiman, V., and Velex, P., 2006, “A Hybrid 3D Finite Element/Lumped Parameter Model for Quasi-Static and Dynamic Analyses of Planetary/Epicyclic Gear Sets,” Mechanism and Machine Theory, 41, pp. 725-748.

[1.35] Abousleiman, V., Velex, P., and Becquerelle, S., 2007, “Modeling of Spur and Helical gear Planetary Drives with Flexible Ring Gears Planet Carriers,” ASME Journal of Mechanical Design, 129, pp. 95-106.

[1.36] Thomas, J., 1991, “A Procedure for Predicting the Load Distribution and Transmission Error Characteristics of Double Helical Gears,” MS Thesis, The Ohio State University.

[1.37] Clapper, M. and Houser, D., 1993, “Prediction of Fully Reversed Stresses at the Base of the Root in Spur and Double Helical Gears in a Split Torque Helicopter Transmission,” Proceedings of American Helicopter Society Rotor Wing Specialists Meeting, Williamsburg, VA.

18

[1.38] Zhang, T., Kohler, H., and Lack, G., 1994, “Noise Optimization of a Double Helical Parallel shaft Gearbox,” International Gearing Conference, UK, pp. 93- 98.

[1.39] Wang, C., Fang, Z., and Jia, H., 2010, “Investigation of Design Modification for Double Helical Gears Reducing Vibration and Noise,” Journal of Marine Science and Applications, 9, pp. 81-86.

[1.40] Jauregui, J. and Gonzalez, O., 1999, “Modeling Axial Vibrations in Herringbone Gears,” Proceedings of ASME Design Engineering Technical Conference, Nevada, DETC99/VIB-8109.

[1.41] Ajmi, M. and Velex, P., 2001, “A Model for Simulating the Quasi-Static and Dynamic Behavior of Double Helical Gears,” The JSME International Conference on Motion and Power Transmission, MPT-2001, pp. 132-137.

[1.42] Anderson, N., Nightingale, L., and Wagner, A., 1989, “Design and Test of Turbofan Gear System,” Journal of Propulsion, 5(1), pp. 95-102.

19

CHAPTER 2

A Linear Time-invariant Dynamic Model of a Double-Helical Planetary Gear Set

2.1 Introduction

In this chapter, a three-dimensional discrete Linear Time-invariant (LTI) model of a

double-helical planetary gear set will be proposed. The modeling methodology and

governing assumptions will be stated. Equations of motion will be derived and solved

using Modal Summation technique in order to predict the steady-state response. The

proposed model will be used to investigate both free and forced vibration characteristics

of an example double-helical planetary gear set within ranges of several key design

parameters. The model will be formulated in a general and modular form such that any

number of equally and unequally positioned planets, any planet-to-planet phasing

conditions as well as any typical support conditions can be simulated effectively.

Furthermore, the proposed model formulation will form the basis for further studies in

Chapters 3 and 4 on gyroscopic and time-varying effects.

20

2.2 Discrete Model and its Assumptions

A three-dimensional discrete model of an N-planet double-helical planetary gear set

is proposed. The formulation adapts 6(N  3) degree-of-freedom (DOF) model of

Kahraman [2.1] to represent the right and left sides of the double-helical planetary gear

independently. It then uses a method proposed by Ajmi and Velex [2.2] to connect the

right and left sides to each other.

The following assumptions are made in the model formulation:

(i) The bodies representing the sun, planet and ring gears and the carrier are all

assumed to be rigid.

(ii) Flexibilities of gear mesh are represented by linear springs acting on the plane

of action normal to gear tooth surfaces (inclined by helix angle ).

(iii) Time-varying component of mesh stiffness due to fluctuation of number of

tooth pairs in contact are neglected in line with the findings of several helical

gear dynamics studies [2.3, 2.4]. Validity of this assumption will be

investigated later in Chapter 4 through a time-varying formulation.

(iv) Gear teeth at the mesh interfaces are assumed to maintain contact all the time

(i.e. tooth separations do not occur). The time-varying formulation of Chapter

4 will also include a clearance type separation function to explore whether

such nonlinear effects are important.

21

(v) The model in this chapter does not include any gyroscopic effects while they

will be investigated in detail in Chapter 3.

(vi) Planets are assumed to be identical to each other such that each planet-sun and

planet-ring mesh can be assumed to have the same geometric properties and

contact characteristics.

(vii) Frictional forces arising from tooth sliding are considered to be negligible in

accordance with the off-line-of-action helical gear vibration measurements of

Kang and Kahraman [2.5].

(viii) Left and right sides of double-helical gears are assumed identical in geometry

except the hands of the teeth are opposite. Right-to-left stagger angles are

assumed to be exactly the same for each corresponding mesh. Any stagger

deviations due to manufacturing will be ignored.

(ix) A class of potential manufacturing errors associated with the gears and the

carrier will be neglected. Such errors including gear run-out and tooth

indexing errors, planet tooth thickness errors, carrier eccentricity, planet pin

hole position errors and ring gear roundness error would impact the dynamic

response in two ways. First of all, each of the run-out and eccentricity errors

would constitute low frequency excitations to be included in the dynamic

model. Secondly, many of these errors prevent an equal load sharing amongst

the planets such that the gear mesh frequency excitation and mesh stiffness of

22

each planet mesh would differ. As the intended applications will be high-

precision aerospace gearing, these errors are of secondary importance.

(x) Damping of the system is represented by either a constant modal damping or a

proportional damping matrix.

Figure 2.1 shows the overall dynamic model of an entire double-helical planetary gear set

with only one planet shown for simplicity purposes. Under the assumptions listed above,

the model formulations will be done first for three basic sub-systems: (i) a sun-planet i

pair (left or right side), (ii) a ring-planet i pair (left or right side), and (iii) a carrier-planet

i pair (left or right side). A beam formulation will then be introduced to combine left and

right sides of each double-helical gear. A general assembly process will then be

employed to obtain the overall mass and stiffness matrices including support bearing

conditions.

2.2.1 A Sun-Planet i Pair Formulation

Figure 2.2 illustrate a dynamic model of an external helical gear pair, which

represents one side (either left or right side) of the sun gear (subscript s) meshing with the

same side of planet-i (subscript pi) located at arbitrary position angle pi . As shown in

Figure 2.2, the plane of action of the gear pair makes an angle  with the vertical y spi

axis. This angle can be defined in terms of the transverse pressure angle  of the sun- sp

planet pair and  as pi

23

y ring gear

planet i

x

sun gear left side

right side z

Figure 2.1: Dynamic model of a double-helical planetary gear system.

24

ypi

ypi spi planet i

ys xpi xpi  ys etspi ()

ksp zpi  pi

zpi x s xs

zs

zs

Sun gear

Figure 2.2: Dynamic model of sun-planet i pair.

25

sp pi, T s : Counterclockwise, spi  (2.1) sp   pi, T s : Clockwise,

where Ts is external torque acting on the sun gear. The undamped equations of motion for this s-pi pair are derived using the helical gear pair formulations of Kahraman [2.1].

These equations for sun gear degrees of freedom are

mys s() t k sp cos cos spi p spi () t 0, (2.2a)

mx() t k cossin p () t 0, (2.2b) s s sp spi spi

mz() t k sin p () t  0, (2.2c) s s sp spi

Itkr() sin  cos pt ()  0, (2.2d) sys sps spispi

Itkr() sin  sin pt ()  0, (2.2e) sxs sps spispi

JtkrptTN() cos  ()  /(2 ). (2.2f) szs sps spi s

The corresponding equations of motion for the degrees of freedom of planet i are

myppi () t k sp coscos spispi p () t  0, (2.3a)

mx() t k cossin p () t 0, (2.3b) ppi sp spispi

mz() t k sin p () t  0, (2.3c) ppi sp spi

Itkr() sin  cos pt ()  0, (2.3d) p ypi sp p spi spi

Itkr() sin  sin pt ()  0, (2.3e) p xpi sp p spi spi

Jtkrpt() cos  ()  0. (2.3f) p zpi sp p spi 26

In these equations, m , I and J are mass, the diametral mass moment of inertia and   

the polar mass moment of inertia of one side (left or right), and r is base circle radius of 

gear  ( spi, ). ksp is the average value of the gear mesh stiffness for the sun-planet

pair. p ()t represents the relative mesh displacement of the s-pi mesh in the direction spi

normal to tooth surface such that

 ptspi() ( yy s pi )cos  spi ( xx s pi )sin  spi r s zs r p zpi cos  + (rrsys p ypi )cos  spi  ( rr s  xs p xpi )sin  spi  zz s  pi sin  et spi ( ) (2.4)

where etspi () is the loaded static transmission error excitation at the s-pi mesh and  is the helix angle. Equations (2.2) and (2.3) are written in matrix form as

11 12 M0q ()ttt KK  q () ff () ss spispissmsi   ksp  . (2.5a) 0M q ()ttt22 q () f () ppisym. Kspi  pi spi  where

yts ()  ytpi ()    xs ()t  xpi ()t     zts ()  ztpi () qs ()t   , q pi ()t    , (2.5b,c) ys ()t ypi ()t   ()t  ()t  xs  xpi     zs ()t zpi ()t

Ms  DiagmmmI ssssss I J, (2.5d) 27

 M ppppppp Diag m m m I I J , (2.5e)

 cc22  csc  2   ccs rccsrcscs 2   rcc  2    ss s 22 2 2  s c   s cs rcsss  cs rs  cs rsc s      222 11  srcsrssrcsss s (2.5f) K   , spi 22 2 2 2 2  rcsrcssrccsss  s  22 2 2   Sym. rss s s r s c s   22   rcs  

 c22  c   c s c 2  c  cs rc 2  cs rc  s cs rc  c 2    pp p  csc222   s  c  scs  rcscs  rs 2  cs rsc  2    pp p  222  ccs scs   s   rcspp    rss    rcs p  K12    , spi  22222  rcss  c s rc  s c s rc sspspsp  s  rr c  s  rr c  s s  rr c  c s   22222  rcscssssspspsp  rs  cs rss   rrcss   rrs  s  rrscs   22 2  rcsssspspsp c  rs c rc  s rr c  c s rr s  c s rr c  (2.5g)

 c22 c  c  s c 2   c cs  rc 2  cs  rc  s cs  rc  c 2    pp p  s22 c   s cs  rc  s cs  rs 2  cs  rs  c 2    pp p  222  srcsrssrcspp p K22    , spi  22 2 2 2 2   rpp c s r css  r p ccs   22 2 2   Sym. rpp s s r s  c s   22   rcp   (2.5h)

28

cc  cc   cs cs     s  s  fsi ()tket sp spi () , fspi ()tket sp spi ()rsc  , (2.5i,j) rss  c  p  rs s rssp   s   rc s rcp  

0  0 0 fsm  . (2.5k) 0 0  TNs /2

In the above equations (2.5f-k), c cosspi , s sinspi , c cos , and ssin  .

2.2.2 A Ring-Planet i Pair Formulation

Next consider the same planet i meshing with the ring gear (subscript r) on one side of the double-helical gear set as shown in Figure 2.3. With pi being at the same angular position pi , the plane of action of this internal gear mesh makes an angle rpi with the vertical y axis, which is defined in terms of the transverse pressure angle rp of the ring-planet pair and  pi as

rp pi, T s : Counterclockwise, rpi  (2.6) rp   pi, T s : Clockwise

29

y pi

ring gear  pi planet i rpi

yr xpi xpi

yr x etrpi () r

 pi  zpi krp xr zpi

zr  zr 

Figure 2.3: Dynamic model of a ring-planet i pair.

30

The undamped equations of motion for the r-pi pair shown in Figure 2.3 are derived for the motions of the ring as

myr r() t k rp coscos rpi p rpi () t 0, (2.7a)

mxrr () t k rp cossin rpirpi p () t 0, (2.7b)

mz() t k sin p () t  0, (2.7c) r r rp rpi

Itkrryr() rpr sin  cos rpirpi pt ()  0, (2.7d)

Itkr() sin  sin pt ()  0, (2.7e) rxr rpr rpirpi

JtkrptTNrzr() rpr cos  rpi ()  r /(2 ). (2.7f)

Here m , I and J are mass, the diametral mass moment of inertia and the polar mass r r r moment of inertia of one side of the ring gear, and r is base circle radius of the ring r gear. The corresponding motions of planet i are governed by

myppi () t k rp cos cos rpirpi p () t  0, (2.8a)

mx() t k cossin p () t  0, (2.8b) ppi rp rpirpi

mzppi () t k rp sin p rpi () t 0, (2.8c)

Itkrp ypi() rp p sin  cos rpi pt rpi ()  0, (2.8d)

Itkrp xpi() rp p sin  sin rpi pt rpi ()  0, (2.8e)

Jtkrptp zpi() rp p cos  rpi ()  0. (2.8f)

31

In Eq. (2.7) and (2.8), krp is the average gear mesh stiffness of the ring-planet pair, and p ()t is the relative displacement of the r-pi mesh along the plane of action normal to rpi the tooth surfaces, defined as

 ptrpi() ( yy r pi )cos  rpi ( x pi x r )sin  rpi r r zr r p zpi cos  +  (rrp ypi r yr )cos  rpi ( rr r xr p xpi )sin  rpi zzet r pi sin  rpi ( ) (2.9)

with etrpi () being the loaded static transmission error excitation at the r-pi mesh. After substituting Eq. (2.9) in, equations (2.7) and (2.8) can be put into the matrix form as

11 12 M0q ()ttt KK  q () ff () rr rpirpirrmri   krp  . (2.10a) 0M q ()ttt22 q () f () ppisym. Krpi  pi rpi 

where q pi ()t is defined in Eq. (2.5c),

ytr ()  xr ()t  ztr () qr ()t   , (2.10b) yr ()t  ()t xr  zr ()t

Mr Diag m rrrrrr m m I I J , (2.10c)

32

 c22 c  csc 2 ccs   rc 2  cs rcscs  rcc  2   rr r 22 2 2  s c   s cs rcsrrr  cs  rs  cs  rsc      222 11  srcsrssrcsrr r K    , rpi 22 2 2 2 2  rcrr s rcss  rccs r  22 2 2   Sym. rrr s s r s  c s   22   rcr   (2.10d)

 c22 c c  s c 2  c cs rc 2  cs rc  s cs rc  c 2   pp p  csc222   s  c  scs   rcscs  rs 2  cs rsc  2    pp p  222  c cs s  cs  s  rcppp  s   rs  s   rcs  K12    , rpi  22222  rcrr c s  rc  s c s rc rrprprp  s   rr c  s  rr c  s s  rr c  c s   22222 rcscsrr  rs  cs  rss rrprprp   rrcss    rrs  s   rrscs   22 2  rcrr c rs  c  rc rrprprp s rr c  c s rr s  c s rr c  (2.10e)

 cc22   csc 2  ccs   rccsrcscs 2   rcc  2    pp p  s22 c   s cs rc  s cs  rs 2  cs  rs  c 2    ppp  222  srcsrssrcs pp   p  K22    , rpi  22 2 2 2 2   rpp c s r css  r p ccs   22 2 2   Sym. rpp s s r s  c s   22   rcp   (2.10f)

33

cc  cc    cs  cs     s  s  fri ()tket rp rpi () , frpi ()tket rp rpi ()rsc . (2.10g,h) rsr  c  p  rs s  rssp    r   rc r  rcp  

0  0 0 frm  . (2.10i) 0 0  TNr /2

In equations (2.10d-i), ccos  , s sin . c cos rpi and s sinrpi .

2.2.3 A Carrier-Planet i Pair Formulation

Figure 2.4 shows a one side of a carrier-planet i pair, with planet i positioned at the same angle  pi and attached to the carrier via a bearing. Rotation center of the planet i represented by the z axis is at a distance r from the rotational axis z of the carrier. pi c c

The planet bearing that is modeled as a diagonal stiffness matrix

 Kbp Diag k x k y k z k x k y 0 couples planet pi to the carrier c along a circle of radius rc . Bearing forces and moments acting on planet pi due to any arbitrary motions of pi and c of this particular side are defined as

34

carrier

planet i

Figure 2.4: Dynamic model of carrier-planet i pair.

35

Ftyycpiczcpi() k [( y y ) r cos ], (2.11a)

Ftx () kxc [( x x piczcpi ) r sin ], (2.11b)

Ftzzcpicycpicxcpi()[()cossin], k z z r r (2.11c)

Mt() k ( ), (2.11d) y y yc ypi

Mtx () kxxcxpi ( ). (2.11e)

Here, pi can rotate in the z direction with no resistance such that Mtz () 0. With these bearing forces and moments defined, the equations of motion of the c-pi pair are written as

mycc () t F y () t 0, (2.12a)

mxcc () t F x () t 0, (2.12b)

mzcc () t F z () t 0, (2.12c)

ItMtrcyc() y ()  c cos  piz Ft ()  0, (2.12d)

ItMtrcxc() x ()  c sin  piz Ft ()  0, (2.12e)

Jtrczc() c sin  pix Ftr ()  c cos  piy FtTN ()  c /(2 ), (2.12f)

myppi () t F y () t 0, (2.13a)

mx() t F () t 0, (2.13b) ppi x

mz() t F () t 0, (2.13c) ppi z

ItMt() ()  0, (2.13d) pypi y 36

ItMt() ()  0. (2.13e) pxpi x

In matrix form, Eq. (2.12) and (2.13) reduce to

M0q()tt KK11 12 q () cccpicpic  fcm   . (2.14a) 0Mppi q ()tt22 q pi () 0  sym. Kcpi 

where q pi ()t is defined in Eq. (2.5c),

ytc ()  xc ()t  ztc () qc ()t   , (2.14b) yc ()t  ()t xc  zc ()t

Mccccccc Diag m m m I I J , (2.14c)

kkrcyyc00 0 0  kkrs00 0  xxc kkrckrszzczc  0 11  Kcpi  kkrckrcs22 2 0 , (2.14d) yzc zc 22 Sym.0 kxzc k r s 22 22 krsxc krc yc 

37

ky 00000 0k 0000 x 00k 000 K12  z , (2.14e) cpi 00krc k 00 zc y 0000krszc  k x  krcyc krs xc  0000

ky 00000 k 0000 x k 000 K22  z . (2.14f) cpi k 00 y Sym.0 kx  0

0  0 0 fcm  . (2.14g) 0 0  TNc /2

In equations (2.14d-f), ccos  and s sin . pi pi

2.2.4 Coupling of the Left and Right Sides

The sub-system models shown in Figures 2.2 to 2.4 consist of only one side of the double-helical gears. In an actual double-helical system, left and right sides of a gears and the carrier are either one-piece (for planet and sun gears) or connected rigidly (for the

38 ring gear and the carrier). As done earlier by Ajmi and Velex [2.2], left and right sides of gear (s, r, pi) are connected by using Euler type finite beam elements.

In Figure 2.5, a double-helical gear is divided into three pieces: the left side gear segment at an inner diameter of Db , the right side gear segment at the same inner diameter, and a connecting structure of outside diameter Dg . The connecting structure spans from gear face mid-point of the right side to the gear face mid-point of the left side.

Here, partitioning is done such that the total mass of the one-piece gear equals the sum of individual masses of the left and right sides and the connecting structure. The same is true for the inertias as well. The ring gear and carrier connecting structures are also handled the same way by using a two-element finite element model, thus representing each gear by three nodes. The node in middle can be used to connect the gear to the support structure through a stiffness matrix that represents a bearing or spline support.

With this the stiffness and mass sub-matrices for the connecting structures are given as

KK11 12 0 ee11 KKKK22 11 12 , (2.15a) eeee12 2 sym. K22 e2

MM11 12 0 ee11 MMMM22 11 12 (2.15b) eeee12 2 sym. M22 e2 where sr,, pic , and subscript e represents beam elements for each component. 39

(a)

(b)

left side right side gear gear segment segment

node L node M node R

beam beam element 1 element 2

Figure 2.5: (a) Geometry of a double helical external gear, and (b) three-piece model of the double helical gear.

40

The sub-matrices in Eq. (2.15) are defined for the n-th ( n[1, 2] ) beam element as

KK11 12 MM11 12  en en  en en  Ken  , Men  , (2.15c,d) sym K22  sym M22  en  en 

The individual elements of Ken and Men are given in Appendix A. These sub- matrices corresponds to the displacement sub-vector

()q L  qqeM () . (2.16)  ()q R where subscripts L, R and M indicate left side, right side and middle nodes of each components respectively. In case where the connecting structure for a gear (sr,, pi),

()q L and ()q R take place of gear displacement vectors in Eq. (2.5), (2.10) and (2.14).

2.2.5 The Overall System Equations

These sub system matrices defined by Eq. (2.5), (2.10) and (2.14) are assembled systematically along with right-to-left coupling matrices defined by Eq. (2.15) to obtain the overall equations of motion of a double-helical planetary gear set consisting of N planets (a total of NNdof 18( 3) degrees of freedom) as

Mq()tt +Cq  () +Kq () tt F (). (2.17)

41 where q()t is the overall displacement vector, M is the mass matrix, F()t is the force vector and K is overall stiffness matrix ( KK mesh K b ) consisting of gear stiffness matrix Kmesh and support stiffness matrix Kb . M and Kmesh are given in the

Appendix B. Vector q()t includes all of the 18(N  3) degrees of freedom

qse()t  qre()t qce ()t q()t  , (2.18) q ()t pe1   q pNe ()t

Forcing vector F()t can be represented as assembly of sub-vectors as

Fs ()t  Fr ()t 0 F()t  , (2.19a) F ()t p1   FpN ()t where

N  N  ffsmsiL () ffrm () ri L  i1  i1    F0s ()t   , F0r ()t    , (2.19b,c) N  N  ff () ff () smsiR  rm ri R  i1 i1

42

(fspi f rpi) L  F0pi ()t   . (2.19d)  (fspi f rpi) R

A support stiffness matrix Kb is defined in the form

Kbbsbrbc Diag  0K 00K 00K 0 0, (2.20a)

and incorporated in Eq. (2.17) to include the matrices for the sun ( Kbs ), ring ( Kbr ), and carrier ( Kbc ) supports. Here, Kbs , Kbr and Kbc are matrices of dimension 6 that are applied to the middle nodes of the respective connecting structures. They are defined as

 Kbyxzyx Diag k k k k k 0, (2.20b) with src,,. Finally, C in Eq. (2.17) is the proportional damping matrix that is given as

CK+M  . (2.21) where  and  are proportionality constants.

2.2.6 Excitations

In Eq. (2.19a), individual forcing vectors forming F()t include transmission error excitations that are given as a part of relative mesh displacements in Eq. (2.4) and (2.9).

There are a total of 4N of such periodic excitations (one for each individual gear mesh)

43 for an N-planet gear set. These excitations have the same fundamental frequency that is equal to the gear mesh (tooth passing) frequency m . Transmission error excitations at the s-pi and r-pi meshes can be computed by using a gear load distribution model [2.6] as well as the average gear mesh stiffnesses ksp and krp . Each of the periodic excitations et() along the meshes of the right and left sides of the gear set have the same spi waveforms at the same harmonic amplitudes since the planets are assumed to be identical. Yet they possess a unique phasing relationship defined by N, planet spacing angles  pi (p1 0 and iN[2, ]), and the numbers of teeth Zs on the sun gear. The same is true for the r-pi excitations et() where N,  (iN[1, ] ), and the numbers of rpi pi teeth Zr on the ring gear define the relative phasing.

Without loss of generality, the s-p1 mesh on the left side of the gear set is chosen here as the reference mesh. The transmission error excitation at this reference mesh is given in Fourier series form as

L ()L etsp1 () eˆspl cos( lt m spl ), (2.22a) l1 where eˆ and  are the amplitude and phase angle of the l-th harmonic of this spl spl excitation as predicted by the gear load distribution analysis [2.6]. The superscript L in parenthesis indicates a left side mesh. The transmission error functions on the other s-pi meshes ( iN[2, ]) on the left side can be defined relative to the reference s-p1 mesh as

44

L ()L etspi () eˆspl cos( lt m spl l spi ), i [2, N ]. (2.22b) l1

Here  is the phase angle between the s-pi mesh on the left side and the reference s-p1 spi mesh [2.1, 2.7]

Zspi , for CW planet rotation, spi  (2.22c) Zspi, for CCW planet rotation.

where Zs is the number of teeth on the sun gear.

An angle  is defined as the phase difference between the left and right side due stg to intentional nominal stagger of the teeth. For the case when right and left side teeth are aligned perfectly (i.e. they are mirror images of each other) as illustrated in Figure 2.6(a),

0 . Meanwhile, for a 50% stagger condition is shown in Figure 2.6(b), stg stg where the tip of a tooth on the left side aligns with a tooth root on the right side. In practice, the stagger condition is a design parameter whose impact on the dynamic response is yet to be described. With the stagger phase angle defined, the excitation on the s-pi mesh of the right side of the gear set is defined as

L ()R etspi () eˆspl cos( lt  m  spl  l spi  l stg ), i  [1, N ]. (2.22d) l1

Meanwhile, the transmission error excitation at the r-p1 mesh on the left side is defined in relation to the reference s-p1 mesh of the left side as

45

(a)

Left Right

(b)

a1

a1

Left Right

Figure 2.6: Illustration of the right-to-left stagger conditions in a double helical gear pair; (a) and (b) .

46

L ()L etrp1 () eˆrpl cos( lt m rpl l rs ), (2.23a) l1 where rs is phase difference between the reference s-p1 mesh and the r-p1 mesh, both on the left side as defined in Ref. [2.7]. eˆ and  are the amplitude and phase angle rpl rpl of the l-th harmonic of the excitation, again predicted by using a gear load distribution model [2.6]. Similarly, excitations on any other r-pi meshes on the left side are given as

L ()L etrpi () eˆrpl cos( lt  m  rpl  l rpi  l rs ), i  [2, N ], (2.23b) l1 where  is the phase difference between the r-pi mesh and the r-p1 mesh on the left rpi side, which is given by

Zrpi, for CW planet rotation, rpi  (2.23c) Zrpi , for CCW planet rotation.

Here Z denotes number of teeth on ring gear. With the same stagger  defined r stg between the left and right sides, excitations of the ring-planet meshes of the right side of the gear set are defined as

L ()R etrpi () eˆrpl cos( lt  m  rpl  l rpi  l rs  l stg ), i  [1, N ]. (2.23d) l1

47

In above equations, the gear mesh frequency m defined above can be determined from kinematic relationships as a function of rotational speeds of the sun and ring gears s and r as [2.8]

ZZsr s/ ( Z s Z r ), fixed ring gear,  msrrsrZZ / ( Z  Z ), fixed sun gear, (2.24)  Zss , fixed carrier.

2.3 Solution Methodology

By setting F0()t  and C0 , Eq. (2.17) is reduced to

Mq()tt + ( Kmesh K b ) q () 0 . (2.25)

2 The corresponding Eigen Value problem KQ MQ (withKKmesh K b ) of this undamped free system is solved to find the undamped natural frequencies  and the

 [1,N ] corresponding mode shapes Q ( dof ).

The steady state response of the LTI system due to the transmission error excitations defined in Section 2.2.6 is obtained using the Modal Summation Technique.

All of these excitations at meshes are, in general, out-of-phase of each other.

Furthermore, each individual harmonic term l of any excitation also has a different phase angle. As such, the response of the gear set to each harmonic term of each gear mesh transmission error excitation must be determined individually using the Modal

48

Summation Technique. These individual responses must then be summed according to the superposition principle to compute the total steady state response.

The Modal Summation Technique makes the use of the expansion theorem in conjunction with the superposition principle to determine the response from  and normalized Q . First, the forcing vector given in Eq. (2.19a) is expressed as a sum of

4N vectors (each representing the excitation at one gear mesh) as

4N FF()tt  k (). (2.26) k 1 where subscript k denotes a particular gear mesh. Explicit expressions for these forcing vectors Fk ()t are given in Appendix C.

Response to an individual forcing vector Fk ()t is obtained by modal summation as

L Ndof qFkkplmplmpl()tk  ( jelt )ˆ cos(  ), (2.27a) l11 where sr, (s if k is representing an external mesh and   r if the gear mesh is an internal one), j 1 and  represents appropriate phasing terms defined by Eq.

(2.22) or (2.23). Fk is the vector of amplitudes of Fk ()t . lm()j is the complex dynamic compliance matrix given by

49

QQ ()j  , (2.27b) lm 222 ()(2)ljl mm   where Q is the -th normalized mode shape.

Modal or proportional type damping can be employed in this formulation. For the case of modal damping,    . When the user assumes a form of damping that is proportional to the mass and stiffness matrices according to Eq. (2.21), the proportional damping ratio  of the -th mode is defined as

2   . (2.28) 2 where  and  are the proportional damping constants. The overall displacement vector is given as sum of steady-state responses to each of individual excitations:

4N qq()tt  k (). (2.29) k 1

With q()t known, relative gear mesh displacements are computed according to Eq.

(2.4) and (2.9), from which dynamic mesh forces at each of the 4N gear meshes can be obtained as

Ftkpt() (), sr , (2.30) pi p pi

50 where Ftpi () and ppi ()t correspond to the same right or left side gear mesh. Likewise the components of the support bearing forces are computed by using user defined support stiffness matrices and the corresponding gear (or carrier) displacements. For practical design purposes, dimensionless dynamic factors KD are defined as the ratio of the maximum gear mesh force to the static gear mesh force FTss (2 rN s ). Dynamic factor for a given mesh is given as

Max F() t pi pi KD 1. (2.31) Fs

2.4 An Example Simulation

A double helical planetary gear set consisting of four equally spaced planets (

N  4 , 0, ,  and 3 ) with a stationary (non-rotating) carrier is used here as an pi 2 2 example gear set. Table 2.1 lists its basic gear design parameters of this example gear set. Total number of degrees of freedom for the model is NNdof 18( 3) 126 .

Individual contact analyses of one sun-planet and one ring-planet mesh were carried out by using a gear pair load distributions model [2.6] under specified loading conditions to determine the average mesh stiffness values ( ksp and krp ), and the harmonic components of the transmission error excitations (amplitudes eˆspl and eˆrpl , and phase angles  and  in equations (2.22) and (2.23)). Only the first three harmonics of spl rpl

51

Table 2.1: Basic design parameters of the example gear system

Sun Planet Ring Carrier

Number of Teeth 47 39 125 -- Normal module (mm) 1.81 1.81 -- Helix angle (o) 21.5 21.5 -- Normal Pressure angle (o) 22.5 22.5 -- Base radius (mm) 41.8 34.7 111.2 -- Mass (kg) 2.4 0.73 6.8 14.4 Mesh Stiffness (N/µm) 564 531 --

kkyx, (N/µm) 100 100 1000 --

kkyx ,   (1e6 Nm/rad) 5 5 10 -- D (mm) g 75 250 64 -- DD/ (mm) -- bo 30 263.8 42  (s) 1.35e-6  (s1 ) 50

kp (N/µm) 564.75 -- 531.15 --

52

Table 2.2: Harmonic amplitudes and phase angles of the transmission error excitations of the example gear set of Table 2.1. [2.6]

Harmonics, eˆ ( µm) eˆ (µm)  (rad)  (rad) l spl rpl spl rpl 1 0.405 0.483 -0.769 -0.709 2 0.088 0.135 -1.299 -1.543 3 0.021 0.005 0.604 0.927

53 the excitations were considered (i.e.l [1, 3] ) as harmonic terms with l  3 have negligibly small amplitudes. Connecting structures between left and right sides have the dimensions listed in Table 2.1 according to Figure 2.5. Table 2.1 also lists the support stiffness values for the sun and ring gears and the carrier as well as the planet bearing stiffness values used in these simulations.

Eigen value solution is carried out for the system defined in Table 2.1 to predict the undamped natural frequencies  as listed in Table 2.3. The corresponding modes are classified as in-phase, sequentially phased and counter-phased, adapting the planetary gear set mode classification of Refs. [2.1, 2.9]. Such classifications that are based on predominantly two-dimensional motions along the transverse plane (x-y plane) of the gears are not fully descriptive here since most of the natural modes exhibit three- dimensional motions with dominant axial (z) and rotational ( x and y ) motions. It is noted in Table 2.3 that there are numerous natural modes within the operating frequency range of the gear set while it is not obvious which modes would be excited by the excitations defined in Section 2.2.

2.4.1 Influence of Right-to-left Stagger

Figures 2.7 shows variation of the maximum dynamic mesh force amplitudes

()FMaxFt  () and ()FMaxFt  () for sun-planet and ring-planet spi L spi L rpi L rpi L meshes of the left side of double helical gears, respectively, with the gear mesh frequency

54

Table 2.3: Predicted natural frequencies and mode types of the example gear set

Mode index Natural Frequency (kHz) Mode Type

  Q

1 0.987 In phase 2, 3 1.262 Sequentially phased 4 1.902 In phase 5,6 1.965 Sequentially phased 7,8 2.076 Sequentially phased 9 2.506 In phase 10 2.518 In phased 11,12 2.522 Sequentially phased 13,14 2.994 Sequentially phased 15 3.166 Sequentially phased    29 8.398 Counter phased    32,33 8.852 Sequentially phased 34 9.037 In phase 35,36 9.451 Sequentially phased    38 11.047 Counter phased   

55

2000 (a) - - - - 1600

1200

800 (N)

400

0 3000 (b)

2400

1800

(N) 1200

600

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 2.7: Maximum dynamic mesh force amplitudes at the left side (a) s- pi and (b) r-pi meshes.

56

2000 (a)

1600 - - - -

1200

(N) 800

400

0 3000 (b) 2400

1800

(N) 1200

600

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 2.8: Maximum dynamic mesh force amplitudes at the right side (a) s-pi and (b) r-pi meshes.

57

m (mssrrZZ   for the case of fixed carrier where s and r are the rotational speeds of the sun and ring gears in rad/s, respectively). Figure 2.8 presents the same for the right side maximum dynamic mesh force amplitudes

()FMaxFt  () and()FMaxFt  () . In both figures, results for three spi R spi R rpi R rpi R right-to-left gear teeth stagger phase angles of 0,  and are compared. Several stg 2  observations can be made from Figures 2.7 and 2.8:

 Maximum dynamic sun-planet mesh force amplitudes at all planet meshes at a

given (left or right) side are equal (i.e. ()()()()FFsp1234 L sp L FF sp L sp L )

and ( ()()()()FFspR1234 spR FF spR spR). The same is true for the ring-

planet mesh forces as well.

 Each of the resonance peaks is associated with a particular mode excited by a

certain harmonic amplitude l of the excitations. The frequencies of these

modes as well as the harmonics exciting these modes are specified as labels of

each resonance peak.

 Maximum dynamic mesh forces vary considerably with stg considered. Not

only amplitudes of response but also the frequencies of resonance peaks

change with stg . Different values of stagger are seen to excite different types

of modes. For stg 0, excited modes exhibit motion where both left and

right sides of double helical gears move together as one piece as in the case of

58

modes at natural frequencies of   1262 and 11047 Hz. These modes are

illustrated in Figure 2.9. Modes 1262 Hz is classified in Table 2.3 as

sequentially phased mode while 11047 which is excited by 2nd harmonic of

excitation is Counter Phased mode. If the modal displacements for

sequentially phased mode   1262 is applied to relative gear mesh

displacement expressions of equations (2.4) and (2.9) with transmission error

terms discarded, one would arrive at templates for this mode in the form

LR LR sp1 a11 a rp1 b11 b sp2 a22 a rp2 b22 b . (2.32) sp3  a11 a rp3  b11 b sp4  a22 a rp4  b22 b

It is noted in these modes that the right and left side gear meshes move in

unison, and hence get excited when stg 0. On the other hand the

sequentially phased modes at   2994 and 3166 Hz exhibit equal but

opposite motions on the right and left sides (as shown in Figure 2.10)

represented by the relative gear mesh displacement templates

LR LR sp1  a11 a rp1 b11 b sp2 a22 a rp2  b22 b (2.33) sp3 a11 a rp3  b11 b sp4  a22 a rp4 b22 b

59

Figure 2.9: Mode shapes representative of Eq. (2.32) at (a) Hz and (b) Hz.

60

Figure 2.10: Mode shape representative of Eq. (2.33) at (a) Hz

and (b) Hz.

61

As a result, these modes are excited when stg . Any stagger angles other

than 0 and  excite both types of motions at lower levels, as illustrated in

response curves for  . stg 2

 For the cases when stg 0 or  maximum gear mesh forces at the left and

right side meshes are equal, i.e. ()()FFspi L spi R and ()()FFrpi L rpi R

while ()()FF and ()()FF for any other  value. spi L spi R rpi L rpi R stg

 The second harmonic components of the excitations for stg excite the

modes that follow the template of Eq. (2.32) with left and right sides moving

8398 in unison. For instance, the resonance peak at 1 Hz is a m 2  2

result of this.

Predicted dynamic factor curves corresponding to Figures 2.7 and 2.8 are shown in

Figures 2.11 and 2.12, respectively. Dynamic factors show maximum dynamic loads are

spi higher than static load carried by the system. In Figure 2.11 dynamic factor KD of 1.28 for the mesh frequency of m  8852 Hz indicates dynamic load is 28% higher than the static load experienced by the corresponding mesh which must be taken into account while designing the gears at that speed.

62

1.3 (a - - - -

1.2

1.1

1 1.5

(b 1.4

1.3

1.2

1.1

1 0 2000 4000 6000 8000 10000

(Hz)

Figure 2.11: Dynamic factors at the left side (a) s-pi and (b) r-pi meshes.

63

1.3 (a - - - -

1.2

1.1

1 1.5

(b 1.4

1.3

1.2

1.1

1 0 2000 4000 6000 8000 10000 (Hz)

Figure 2.12: Dynamic factors at the right side (a) s-pi and (b) r-pi meshes.

64

2.4.2 Influence of Planet Phasing Conditions

Two earlier studies by Kahraman [2.1] and Kahraman and Blankenship [2.10] provided the framework which established a well-structured relationship between the mode shapes excited and the phasing of the excitations as formulated in Section 2.2.

These studies classified the natural modes of a four-planet helical (or spur) planetary gear set (with equally spaced planets) in three groups:

 In-phase (IP) modes: In these axisymmetric modes, all planets move in an

identical manner relative to the central member of the gear set (sun, ring and

carrier). These modes were shown to be excited by in phase harmonic terms

of the excitations.

 Sequentially phased (SP) modes: In these modes, each planet moves in a

certain way relative to the central members such that none of the central

members exhibit any motions. In other words, these modes are limited to

planet motions. Kahraman [2.1] and Kahraman and Blankenship [2.10]

showed that these modes are excited by the harmonics of the excitation that

are sequentially phased.

 Counter phased (CP) modes: As a special case of sequentially phased modes

unique to four-planet gear sets, diametrically opposed planets move the same

way relative to the central members while the motions of two adjacent planets

are 180-degrees out-of-phase. The central members do not move in these

65

modes. The same studies also indicated that these modes are excited by

counter-phased harmonics of the excitations.

These above rules were only valid for gear sets consisting of a single flank (spur or single helical). Therefore, they might not be fully applicable to the double-helical arrangement in hand. In order to investigate the influence of such planet phasing conditions, three variations of the example gear set are considered

 A Sequentially Phased Gear Set: The example gear set specified in Table 2.1

has number of teeth values of Z  47, Z 125 and Z  39. As s r p

ZNs 47 4 11.75 and ZNr 125 4 31.25, the following phase angles

apply in Eq. (2.2):

For l  1: lZ 0, ,  , 3 for i 1, 2, 3, 4. spi 22 lZ  0, , 0,  i 1, 2, 3, 4. For l  2 : spi for

For l  3 : lZ 0,3 ,  ,  for i 1, 2, 3, 4. spi 22

For l  4 : lZspi  0, 0, 0, 0 for i 1, 2, 3, 4.

A very similar phasing relationship applies to the excitations of the ring gear

meshes as well. These excitation phasing conditions indicate that the first and

third harmonics of the excitations of this gear set should excite sequentially

phased modes while the second and fourth harmonic should excite counter-

phased and in-phase modes, respectively.

66

 An In-phase Gear Set: A gear set with Z  48, Z 124 and Z  39 has s r p

ZNs 48 4 12 and ZNr 124 4 31 with lZspi0, 0, 0, 0 for

i 1, 2, 3, 4 regardless of the value of l. The same is true for the ring gear meshes as well. This suggests that all harmonics of the excitation should

excite the in-phase modes only.

Z 46, Z 126  A Counter-phased Gear Set: A third variation with s  r  and

Z p  39 has ZNs  46 4 11.5 and ZNr 126 4 31.5 with the

resultant phase angles:

For l 1, 3 : lZspi 0, , 0,  for i 1, 2, 3, 4.

l  2, 4 lZ   0, 0, 0, 0 i 1, 2, 3, 4. For : spi for

These excitation phasing conditions indicate that the first and third harmonics

of the excitations of this gear set should excite counter-phased modes while

the second and fourth harmonic should excite in-phase modes.

In Figure 2.13(a), the maximum s-pi gear mesh forces of these three gear sets are compared for the stagger condition of stg 0 while Figure 2.13(b) shows the same for

stg . For stg 0 , the SP gear set is seen to excite SP modes while the IP gear set exciting IP modes and the CP gear set being concerned with the CP modes. These SP, IP and CP modes follow the following templates for relative mesh displacements, respectively:

67

LR LR LR

sp1 a11 a sp1 a11 a sp1 a11 a sp2 a22 a sp2 a11 a sp2  a11 a (2.34a-c) sp3  a11 a sp3 a11 a sp3 a11 a sp4  a22 a sp4 a11 a sp4  a11 a

It is noted here that right and left side motions are in unison in these excited modes regardless of the planet phasing conditions since stg 0 .

For a 50% stagger ( stg ) in Figure 2.13(b), the phasing between the right and left sides of the double helical gear set become evident with the types of modes compared in Figure 2.13(a). The SP, IP and CP modes excited by the corresponding gear sets have the following respective relative mesh displacement templates:

LR LR LR

sp1  a11 a sp1  a11 a sp1  a11 a sp2 a22 a sp2  a11 a sp2 a11 a . (2.35a-c) sp3 a11 a sp3  a11 a sp3  a11 a sp4  a22 a sp4  a11 a sp4 a11 a

All three of templates above, of these types of modes point to motions that are 180- degrees out-of-phase between the right and left sides of the double-helical gear set.

These are the modes that require a 50% stagger in order for them to be energized.

2.4.3 Influence of Number of Planets

In order to study the effect of number of planet gears N of the dynamic response of the gear set, certain basic design parameters of the example gear set are modified. Three 68

3600 (a) - - - -

2700

1800 (N)

900

0 2000

(b)

1500

1000

(N)

500

0

(Hz)

Figure 2.13: Maximum dynamic mesh force amplitudes at the left/right side s- pi meshes for (a) and (b) for different planet

phasing conditions.

69

1600

(a) - - - -

1200

800 (N)

400

0 1600

(b)

1200

800 (N)

400

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 2.14: Maximum dynamic mesh force amplitudes at the left/right side (a) s-pi and (b) r-pi meshes for different number of planet gears.

70 to five planet variations of the same example gear set are devised all having Zs  53,

Zr 127 and Z p  37 . This number of teeth combination allows the three, four and five-planet gear sets to have equally spaced planets as well as all three gear sets having sequential phasing conditions. Figure 2.14 compares the dynamic responses of these three gear set with stg  0. Considerable changes in dynamic response can be observed under same loading and damping conditions for different N. While the three-planet gear set exhibits the largest resonance peaks, there are no clear trends on which gear set would be better dynamically regardless of the operating speed ranges and torques.

2.4.4 Influence of Radially Floating Sun Gear

In planetary gear sets, the sun gear is often allowed to float radially (i.e. not supported or piloted radially by a bearing support) in order to allow the gear set to “self- center” itself to compensate for certain types of carrier and gear manufacturing errors as explained by and Ligata and Kahraman [2.11] and Bodas and Kahraman [2.12]. In Figure

2.15, the influence of floating the sun gear radially on the forced response curves is demonstrated for both stg  0 and  on the baseline design of Table 2.1. In case of a floating sun, diagonal terms of the sun gear support stiffness matrix K in Eq. (2.20b) bs are taken to be kk1(10)6 N/m, kk5(10)4 Nm/rad while the piloted ys xs ys xs gear set uses the values given in Table 2.1. Modes of the form given by Eq. (2.33) appear to be most sensitive to the values of the sun gear radial support stiffnesses. As a result,

71

2000 Piloted Sun (a) - - - - Floating sun 1500

1000 (N)

500

0 2000 (b)

1500

(N) 1000

500

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 2.15: Effect of radially floating sun gear on dynamic mesh force amplitudes for (a) and (b) .

72 the response of the gear set with the stagger condition of   shown in Figure stg

2.15(b) exhibits significant changes at lower frequency ranges (say m 5000 Hz) when the sun gear is radially floating. However very little influence of sun gear support is evident in Figure 2.15(a) for stg 0 .

2.5 Mode Identification using Modal Strain Energy

The Eigen value analysis of the undamped free system yields natural frequencies

 Q  and corresponding mode shapes  (  [1,Ndof ] ) as described in Section 2.3. The forced response curves shown in Figures 2.7, 2-8, 2.13-2.15 indicate that only a few of these modes are excited by the transmission error excitations applied at the gear meshes.

Unless a forced response computation is performed, the question of which modes should be expected to be excited is left unanswered. In an attempt to answer this question, modal strain energies of each mode will be formulated and computed in this section.

The same strain energy computation should also identify the most heavily loaded component and corresponding degrees of freedom within a given mode shape.

Modal strain energy U of the -th vibration mode can be expressed as

1 U  QKQT . (2.36a) 2

73 where K is overall stiffness matrix as given in Eq. (2.17). As a modal quantity, the value of U has no meaning since the mode shapes are given in terms of modal displacements that are relative quantities as

qˆ se ˆ qre qˆ ce Q  , (2.36b)  qˆ pe1   ˆ q pNe

where qˆe ( srcpi,,, .) can be expressed as

()qˆ  L  qqˆˆeM () . (2.36c)  ()qˆ  R

Using these modal vectors, total modal strain energy can be grouped in three categories

[2.13] such that

UU sup U mesh  U pb. (2.37)

Here U represents the total modal strain energy in support bearings for sun and ring sup

U gears and the carrier, mesh represents the total strain energy in all gear meshes and

U denotes the total strain energy associated with the planet bearings. pb

Strain energy in support springs U is defined as sup 74

UUUUsup s r c , (2.38a)

where Us , Ur and Uc are the strain energies in support springs of the sun, ring and

carrier, respectively. These strain energies components are defined as

11 Ukykxkzkk()qKqˆˆT () ˆˆ222ˆ ˆˆ 2 2 , (2.38b) 22Mb M y x z yy  xx where K represents the support stiffness matrix for sun, ring and carrier ( src,, ) b

applied to middle nodes of the respective connecting structures as given by Eq. (2.20b)

and

yˆ  xˆ   zˆ ()qˆ   . (2.38c)  M ˆ y  ˆ x  ˆ z

is the modal displacement vector of the corresponding middle node. This strain energy

can be further broken down into its translational, axial and tilting components by

considering appropriate degrees of freedom.

U Strain energy in gear meshes mesh can be divided into sun-planet and ring-planet

meshes as follows

75

N UUUUU   UUUU    . mesh sp rpLR sp rp  spi rpi L spi rpi R i1 (2.39) where U and U are the strain energies in sun-planet i and ring-planet i meshes, spi rpi respectively, and subscripts R and L denote right and left sides. These strain energies can be expressed in terms of modal relative mesh displacements pˆ for sun gear meshes spi and pˆ for ring gear meshes. They are defined from Eq. (2.4) and (2.9) as rpi

ˆˆˆ ˆˆ ˆˆ pyyspi()cos()sin s pi  spi xx s pi  spi rr s zs p zpi cos (2.40a) ˆˆ ˆˆ ˆˆ + (rrs ys p ypi )cos  spi  ( rr s  xs p xpi )sin  spi  zz s  pi sin 

ˆˆˆ ˆˆ ˆˆ pyyrpi()cos()sin r pi  rpi xx pi r  rpi rr r zr p zpi cos (2.40b) ˆˆ ˆˆ ˆˆ +  (rrp ypi r yr )cos  rpi ( rr r xr p xpi )sin  rpi zz r pi sin 

With these, U and U are given as spi rpi

1 2 Ukp ˆ , (2.41a) spi2 sp spi

1 2 Ukp ˆ . (2.41b) rpi2 rp rpi

Finally the modal strain energy in planet bearings involves the relative modal displacements between the carrier and planets on left and right sides such that

76

1 N 22  ˆˆˆˆ ˆˆ Ukyyrkxxrpb  y c pi c zccos pi x c pi c zc sin pi 2 i1 2 ˆˆ ˆˆ kzz c z pi r c yccos pi r c xc sin pi (2.42) ˆˆ22 ˆˆ  kky yc ypi x xc xpi .  which can be further separated into its translational, axial and tilting components.

As stated earlier, the absolute value of U in Eq. (2.36a) has no meaning since the modal displacements in Eq. (2.36b) are relative quantities. For this reason, the value of each strain energy component can be normalized such that U 1.

Strain energy distribution for different modes of the example gear set is given in the

Table 2.4. The modes excited by the transmission error are presented in italic characters in this table. Most of the excited modes in the higher frequency region (say  3000

Hz) exhibit higher strain energy content in gear meshes as compared to strain energies in

8398 support structures and planet bearings. The excited modes at frequencies 29 and

35,36 9451Hz shows normalized Umesh values of 0.89 and 0.93, respectively. This indicates that most of the energy is represented by the relative gear mesh displacements in these modes, making them the most likely modes excitable by the transmission error excitations. Further break down of strain energies at these modes is shown in Figure 2.16.

8398 The mode at 29 Hz exhibits nearly same level of energy in the sun-planet and ring-planet meshes, while the mode at 35,36  9451 Hz shows more strain energy in the ring-planet meshes. 77

Table 2.4: Strain Energy distribution for the modes of the example gear set (excited modes are shown in italic characters)

Mode Index Natural Frequency (kHz) Modal Strain Energy

  Umesh Usup U pb 1 0.987 0.07 0.00 0.93 2, 3 1.262 0.02 0.62 0.36 4 1.902 0.93 0.00 0.07 5,6 1.965 0.00 0.83 0.17 7,8 2.076 0.18 0.82 0.00 9 2.506 0.00 0.00 1.00 10 2.518 0.00 0.00 1.00 11,12 2.522 0.00 0.01 0.99 13,14 2.994 0.35 0.63 0.02 15 3.166 0.98 0.00 0.02   29 8.398 0.89 0.00 0.11   32,33 8.852 0.91 0.00 0.09 34 9.037 0.89 0.00 0.10 35,36 9.451 0.93 0.00 0.06   38 11.047 0.99 0.00 0.01   

78

0.6

0.4

0.2 Normalized strain energy strain Normalized

0 0.6

0.4

0.2 Normalized strain energy Normalized

0

Figure 2.16: Normalized modal strain energy components of modes at (a) Hz (b) Hz.

79

2.6 Summary

In this chapter, a linear, time-invariant model of a double-helical planetary gear set was developed which allows the analysis of a gear set with any number of planets, any planet phasing and spacing configurations and any support conditions. The model included all rigid body degrees of freedom of gears and the carrier in a three-dimensional manner. Planets were allowed to be positioned equally or unequally spaced around the sun gear. The model captures the phasing relationships between the planet meshes as well as the right-to-left phase differences associated with the staggering the teeth of gears.

Free and forced vibration analyses of the model were carried out by solving the governing Eigen Value problem and applying the modal summation technique with proportional/modal damping.

An example gear set analyses point to various unique dynamic behaviors of double- helical planetary gear sets. Numerous modes with dominant tilting motions are predicted to indicate that a three-dimensional formulation is must for double-helical planetary gear sets. The results of parametric studies also show that right-to-left stagger conditions and planet phasing conditions are equally critical in defining what modes are excited and what resonance peaks are formed in the forced response of the gear set. Change in number of planet gears in the gear set was also shown to change the dynamic response considerably. Additionally, the influence of the sun support conditions was shown to be more pronounced for the case of 50% stagger. At the end, the modal strain energies associated the gear meshes, planet bearings and support structures of the central members

80 is formulated to demonstrate that most of the modes excited by the transmission error excitations are those with high gear mesh modal strain energies.

References for Chapter 2:

[2.1] Kahraman, A., 1994, “Planetary Gear Train Dynamics,” ASME Journal of Mechanical Design, 116, pp. 713-720.

[2.2] Ajmi, M. and Velex, P., 2001, “A Model for Simulating the Quasi-Static and Dynamic Behavior of Double Helical Gears,” The JSME International Conference on Motion and Power Transmission, MPT-2001, pp. 132-137.

[2.3] Seager, D., 1975, “Conditions for the Neutralization of Excitation by the Teeth in Epicyclic Gearing,” Journal Mechanical Engineering Science, 17(5), pp. 293- 298.

[2.4] Kubur, M., Kahraman, A., Zini, D., and Kienzle, K., 2004, “Dynamic Analysis of a Multi-Shaft Helical Gear Transmission by Finite Elements: Model and Experiment,” ASME Journal of Vibration and Acoustics, 126, pp. 398-406.

[2.5] Kang, M. and Kahraman, A., 2012, “Measurement of Vibratory Motions of Gears Supported by Compliant Shafts,” Mechanical Systems and Signal Processing, 29, pp. 391-403.

[2.6] LDP Gear Load Distribution Program, 2011, Gear and Power Transmission Research Laboratory, The Ohio State University, USA.

[2.7] Parker, R. and Lin, J., 2004, “Mesh Phasing Relationships in Planetary and Epicyclic Gears,” ASME Journal of Mechanical Design, 126, pp. 365-370.

[2.8] Kahraman, A., 1994, “Load Sharing Characteristics of Planetary Transmission,” Mechanism and Machine Theory, 29(8), pp. 1151-1165.

[2.9] Platt, R. and Leopold, R., 1996, “A Study on Helical Gear Planetary Phasing Effects on Transmission Noise,” VDI Berichte, 1230, pp.793-807. 81

[2.10] Kahraman, A. and Blankenship, G.W., 1994, “Planet Mesh Phasing in Epicyclic Gear Sets,” International Gearing Conference, Newcastle upon Tyne, pp. 99-104.

[2.11] Ligata, H., Kahraman, A., and Singh, A., 2009, “Closed-form Planet Load Sharing Formulae for Planetary Gear Sets using Translational Analogy,” Journal of Mechanical Design, 131, 021007-1 to 021007-7.

[2.12] Bodas, A. and Kahraman, A., 2004, “Influence of Carrier and Gear Manufacturing Errors on the Static Load Sharing Behavior of Planetary Gear Sets,” JSME International Journal, Series C, 47(3), 908-915.

[2.13] Lin, J., 2000, “Analytical Investigation of Planetary Gear Dynamics,” PhD Thesis, The Ohio State University.

82

CHAPTER 3

Influence of Gyroscopic Effects on Dynamic Behavior of Double-Helical Planetary

Gear Sets

3.1 Introduction

In this chapter, the three-dimensional discrete Linear Time-invariant (LTI) model proposed in Chapter 2 is expanded to include certain classes of gyroscopic effects to study their potential impact on the dynamic behavior of the double-helical planetary gear sets. Two types of gyroscopic effects might exist in a planetary gear set:

(i) gyroscopic moment due to resistance of a spinning body to its change in plane

of rotation (as per principle of conservation of angular momentum), and

(ii) gyroscopic effect due to a rotating carrier, which introduces additional

centripetal and Coriolis acceleration components.

In most geared turbo fan application, carrier is stationary (non-rotating), often due to its weight, mounting complexities as well as lubrication system implementation issues.

With geared turbofan applications in focus, gyroscopic effects due to rotating carrier are 83 not considered in this chapter. Gyroscopic moments due to skew or tilting are included in the linear time-invariant equations of motion. A complex Eigen value solver is employed to determine the natural modes. A complex modal summation formulation is then used to predict the forced response with gyroscopic effects included. Influences of several key design parameters and operating conditions on the dynamic response of the system due to additional gyroscopic moments are studied at the end.

3.2 Incorporation of Gyroscopic Moments in the Dynamic Model

Principle of conservation of angular momentum indicates that a spinning body resists any force tending to change its plane of rotation. Resistance to this change depends on mass moment of inertia of the spinning body and its angular rotational velocity. Accordingly, any tilting/rocking motion of a gears or carrier is resisted by a gyroscopic couple (moment). A generalized formulation for this gyroscopic moment is derived here for any spinning body  (  sr,, pi), rotating at constant angular velocity of k where k is the unit vector along the z axis. The angular momentum of the spinning body with no tilting motion (rotating in its plane of rotation) is

HkJ , (3.1)

The velocity vector for the same body due to tilting and spinning motions can be expressed as

xyi+  j  z  k, (3.2)

84 where i, j and k are unit vectors along x, y and z axis respectively and  x ,  y and

 z are vibratory velocities in x, y and z direction respectively. As per principle of conservation of angular momentum, rate of change of angular momentum due to tilting motion leads to additional moment M given by

MH  H, (3.3a)

Substituting Eq. (3.1) and (3.2) in Eq. (3.3a) gives expression for additional moment acting on body  as

MijJJyx. (3.3b)

The components of this additional moment in x and y direction (i and j components respectively) can be incorporated in linear time invariant equations of motion as derived in chapter 2. Using Eq. (3.3) the equations of motion for each of the sub systems as described in section 2.1 are modified as follows

3.2.1 A Sun-Planet i Pair with Gyroscopic Effects

Equations (2.2) and (2.3) derived earlier for a sun-planet pair shown in Figure 2.2 is modified to account for gyroscopic moments, defined by Eq. (3.3). Due to the vibratory displacements xs ()t and ys ()t of the sun gear, gyroscopic moments are created to modify the sun gear equations of motion in  and  direction (Eq. (2.2d,e) as xs ys

85

ItJsys() s sxs  () tkr sps sin  cos spispi pt ()  0, (3.4a)

ItJsxs() s sys  () tkr sps sin  sin spispi pt ()  0. (3.4b) while other sun gear equations of motion 2.2(a-c,f) remain the same.

Similarly, Eq. (2.3a-c,f) for planet motions remain unchanged and Eq. (2.3d,e) are modified to account for gyroscopic moments as

ItJtkrpypi() p pxpi  () spp sin  cos spispi pt ()  0, (3.5a)

ItJp xpi() p p  ypi () tkr sp p sin  sin spi pt spi ()  0. (3.5b)

Eq. (2.2a-c,f), (3.4), Eq. (2.3a-c,f) and (3.5) are written in matrix form as

M0qssss ()tt G0q   ()    0Mppi  q()tt  0G ppi  q  ()

11 12 KKspi spi q s()tt ff sm si ()  k   . (3.6a) sp 22 sym. K qfpi()tt spi ()  spi

Here velocity sub-vectors are given as

ys ()t  y pi ()t     xs ()t  x pi ()t     zts ()  ztpi () q s ()t   , q pi ()t    , (3.6b,c)  ys ()t  ypi ()t   ()t  ()t   xs  xpi      zs ()t  zpi ()t and the sub-matrices of the gyroscopic matrix are found as 86

000000  000000 000000 Gss , (3.6d) 0000Js 0 000J 00 s 000000

000000  000000 000000 G  . (3.6e) pp0000J 0 p 000J p 00  000000

where psspZ Z is the rotational speed of the planet (with a non-rotating carrier).

It is evident from Eq. (3.6) that gyroscopic effects introduce an additional skew- symmetric component to the overall damping matrix that was defined in Chapter 2 by considering a proportional damping mechanism.

3.2.2 A Ring-Planet i Pair with Gyroscopic Effects

In a similar manner, equations of motion for the ring-planet i pair shown in Figure

2.3 are obtained by modifying Eq. (2.7) and (2.8) to account for gyroscopic moments.

The resultant matrix equation for the ring-planet i pair with gyroscopic effects included is as follows

87

M0qrrrr ()tt G0q   ()    0Mppi q()tt  0G ppi  q  ()   KK11 12  q()tt ff ()  rpi rpi r rm ri  krp   . (3.7a) sym. K22 qfpi()tt rpi ()  rpi   where

yr ()t  xr ()t  ztr () q r ()t   , (3.7b)  yr ()t  ()t  xr   zr ()t

000000  000000 000000 Grr . (3.7c) 0000Jr 0 000J 00 r 000000

The rest of the matrices in Eq. (3.7a) are as defined in Eq. (2.10). In Eq. (3.7c), r is the rotational speed of the ring gear that is defined as  Z Z for a stationary carrier. rrss

3.2.3 A Carrier-Planet i Pair with Gyroscopic Effects

The gyroscopic moments associated with a carrier-planet i pair shown in Figure 2.4 is obtained through modifications to Eq. (2.12) and (2.13). Here the gyroscopic effects

88 caused by the planet motion must be accounted for while no gyroscopic moment acts on the stationary carrier (c 0 ). Matrix form of the equations of motion of this pair is as follows

11 12 M0qcc ()tt 00q   ccpicpic () KK  q () t0     .(3.8a) 0Mppi q()tt 0G ppi q  ()22 q pi () t 0    sym. Kcpi  where

yc ()t  xc ()t  ztc () q c ()t   . (3.8b)  yc ()t  ()t  xc   zc ()t

The remaining matrices in Eq. (3.7a) are as in Eq. (2.14).

3.2.4 Coupling Elements with Gyroscopic Effects

With gyroscopic moments for one side of gears are accounted for, gyroscopic matrices for connecting beam elements are defined next. As described in Section 2.2.4 and illustrated in Figure 2.5, each double-helical gear is divided into three pieces with left and right sides connected using Euler type finite beam elements. The gyroscopic matrix for each connecting structure is given as

89

GG011 12 ee11 GGGG22 11 12 , (3.9a) eeee12 2 skew sym. G22 e2 where sr,, pi and subscript e denotes beam elements for each component. The sub- matrices in Eq. (3.9a) are defined for the n-th ( n[1, 2] ) beam element as

11 12 GGen en G  . (3.9b) en 22 skew sym. Gen

The individual elements of Gen are given in Appendix A.

3.2.5 The Overall System Equations with Gyroscopic Effects

The sub-system gyroscopic matrices for gears and carrier as defined by Eq. (3.6) to

(3.8) along with gyroscopic matrices for connecting structures given in Eq. (3.9) are assembled systematically to form equations of motion of the entire double-helical planetary gear set as

Mq()tt + G+C q  () + Kmesh K b  q () tt F (). (3.10)

All of the system matrices here are defined in Eq. (2.17) and Appendix B.

90

3.3 Solution Methodology

A free vibration analysis of a given system is carried out for undamped case (C0

) to study the influence of gyroscopic moments on the natural frequencies of the system.

The response of the linear, time-invariant system to the transmission error excitations is again obtained by using the Modal Summation Technique in complex domain [3.1]. For forced vibration response a proportional damping CK+M   is considered here together with the gyroscopic matrix G . For this purpose, Eq. (3.10) is put in the state- space form

rAr+BF()ttt () (), (3.11a) with the state vector r()t , and matrices A and B defined as

q()t r()t   , (3.11b) q()t

0I  0  A  , B= . (3.11c,d) 11  1 MK M[G+C] M 

Here I is the identity matrix of dimension Ndof . The free vibration solution of Eq.

(3.11a) can be carried out by setting F0()t  which reduces Eq. (3.11a) to

rAr()tt (). (3.12)

The corresponding algebraic and adjoint Eigen value problems for a system are given, respectively, as 91

AR  R , (3.13a)

LATT  L. (3.13b) where  a diagonal matrix of Eigen values and R and L are the modal matrices of so- called right and left Eigenvectors, respectively. The modal matrices R and L are biorthogonal such that LRT  I and LART   .

As explained earlier a free vibration analysis is carried out first for an undamped system (C0 ). The Eigen values here come in complex conjugate pairs. The -th Eigen value  ()j or its complex conjugate  ()j ) obtained from Eq. (3.13) is used to determine the -th natural frequency as

()imag  ( j  )  imag  ( j  ). (3.14)  

The Eigen values and natural frequencies obtained are speed dependent.

Similar to LTI system without gyroscopic effect, a proportional damping matrix is considered for calculating the response of the system to transmission error excitations.

The forcing vector is separated to its components based on their phasing using same methodology presented in Section 2.3,

4N FF()tt  k (), (3.15) k 1

92 with each of the 4N terms represent the excitation caused by a gear mesh. Expansion theorem is used in conjunction with the above biorthogonality relations to find the response of the given system to individual forcing function Fk ()t [3.1]

L 2Ndof ˆ rkp()tk lmplmpl ( j ) eˆ cos( lt ), (3.16a) l11 where sr, , j 1 and  represents appropriate phasing terms defined by Eq.

ˆ ()j (2.22) or (2.23). The term lm is given by

T ˆ LBF k lm()j R . (3.16b) jm 

Here Fk is the vector of amplitudes of Fk ()t , and R and L are the -th pair of

Eigenvectors. Linearity of the model allows superposition principle to be applied to find the total steady-state response as

4N rr()tt  k (). (3.17) k 1

The state vector r()t consists of displacement vector q()t and the velocity vector q()t as given by Eq. (3.11b). With q()t known, relative gear mesh displacements are computed according to Eq. (2.4) and (2.9), from which dynamic mesh forces at each of the 4N gear meshes can be obtained. The dynamic factors as defined by Eq. (2.31) are valid here as well.

93

3.4 An Example Simulation

The example double-helical planetary gear set used in Chapter 2 with parameters listed in Table 2.1 will be used as well. The gyroscopic effects formulated above are included in the analysis of this four-planet gear set ( N  4 ) with a non-rotating carrier.

Complex Eigen value solution was carrier out to predict the undamped natural frequencies  . Influence of different gear parameters on the natural frequencies of the system was studied. Complex modal summation was carried out to quantify the influence of gyroscopic terms on the forced vibration response of the same system with the same transmission error excitations as in Chapter 2 with additional proportional damping.

As the gyroscopic matrix G is speed dependent (i.e. GG () ), modal behavior becomes speed dependent as well. Rotational speed of input member (sun gear) is varied from 0 to 020,000s  rpm ( 0 s 2094 rad/s) and Eigen value problem was solved at various s values within this range and influence of rotational speed on  was quantified. Figure 3.1 shows variation of  with s . Only some of the natural frequencies within the frequency range of interest are seen to vary with s while others remain the same regardless of s . These affected natural frequencies are listed in Table

3.1 together with the strain energy components associated with the corresponding mode shapes. The modes with lower frequencies ( 7  2076 and 14  2994 Hz.) exhibit

94

10000

p2 9500

(Hz) p2

9000 p1

p1

8500 4000

3500

(Hz) 3000

2500

2000 0 5000 10000 15000 20000

(rpm)

Figure 3.1: Variation of certain natural frequencies with the rotational speed due to gyroscopic effects ( ).

95

Table 3.1: Strain energy distribution for the modes exhibiting change in natural frequencies due to gyroscopic effects.

Mode Index Natural Frequency (kHz) Modal Strain Energy

  U U U  mesh sup pb 7, 8 2.076 0.18 0.82 0.00 13, 14 2.994 0.35 0.63 0.02 17, 18 3.757 0.67 0.32 0.02 32, 33 8.852 0.91 0.00 0.09 35, 36 9.451 0.93 0.00 0.06

96 more strain energy in the bearings (support or planet bearings) with most of it caused by the tilting action as shown in Figure 3.2.

All the affected modes are double modes (two modes at the same natural frequency) when gyroscopic effects are not included. When gyroscopic effects are included, they split and diverge from each other with increasing s as observed in Figure 3.1. For one mode, the natural frequency increases due to gyroscopic stiffening while the frequency of the companion mode reduces due to gyroscopic softening. The maximum change in natural frequency in Figure 3.1 is observed for the mode at 17,18  3757 Hz. This frequency value predicted without gyroscopic effects changes to 17 3675 and

18 3887 Hz at s 20,000 rpm, representing a modest 3.4% change.

Maximum percent change in natural frequencies within the same speed range was used as a metric to quantify the influence of support stiffness on the gyroscopic moments.

Figure 3.3(a) shows that the maximum percentage change in natural frequencies due to gyroscopic effects with the sun stiffnesses in tilting/rocking directions. The changes in the natural frequencies is the most significant at the lowest values of k and ys kxs . This is rather predictable since reducing stiffnesses in rocking direction result in larger amplitudes of ys and xs (and hence,  ys and  xs ). Similar effect is observed as a result of the changes in the ring gear bearing stiffnesses k and k in yr xr tilting/rocking directions as well, as shown in Figure 3.3(b).

97

0.8 (a)

0.6

0.4

0.2

0 0.8

(b)

0.6

0.4

0.2

0

Figure 3.2: Strain energy distribution for support spring and planet bearings for modes at (a) Hz and (b) Hz. (Tr - Translational, Ti - Tilting, Ax - Axial).

98

20 (a)

15

Max change 10 in [%]

5

0 012345 (Nm/µrad)

20 (b)

15

Max change 10 in [%]

5

0 012345 (Nm/µrad)

Figure 3.3: Maximum percent change in natural frequencies with an input speed range of 0 to 20,000 rpm as a function of (a) sun gear support stiffness in tilting direction and (b) ring gear support stiffness in tilting direction.

99

40 (a)

30

Max change 20 in [%]

10

0 0.010 0.015 0.020 0.025 0.030 2 (kg-m )

40 (b)

30

Max change 20 in [%]

10

0 0.10 0.15 0.20 0.25 0.30 2 (kg-m )

Figure 3.4: Maximum percent change in natural frequencies with an input speed range of 0 to 20,000 rpm as a function of polar mass moment of inertia of (a) the sun gear and (b) the ring gear.

100

Next, the influence of the polar mass moments of inertia on gyroscopic effects is investigated. Two cases are considered, one with a floating sun gear configuration (

6 4 kkys xs 10 N/m, kkys xs 5(10) Nm/rad) and other with floating ring gear (

7 5 kkyr xr 10 N/m, kkyr xr 10 Nm/rad). Polar mass moments of inertia of the sun and ring gears are varied in Figure 3.4(a) and (b) for the above cases to observe the maximum percentage change in natural frequencies. This figure shows that, as polar mass moments of inertia of gears increase, maximum percent change in natural frequencies with speed increases as well.

Gyroscopic effects can be expected to influence the forced frequency response curves of the gear set in two ways. One is the resonant frequencies assuming the modes listed in Table 3.1 are excited by the transmission error excitations. The other is the changes done to the damping matrix with the addition of G, which should influence the amplitudes of some of the resonance peaks. Dynamic gear mesh force amplitude curves with gyroscopic effects included are compared in Figure 3.5 to those of the LTI system without gyroscopic effects for a case of no right-to-left stagger ( stg 0). Somewhat lower proportional damping values with coefficients (5(10)7 s, 10 s-1) are used here as compared those in Chapter 2. It is seen in Figure 3.5 that there is little difference between the curves with and without gyroscopic effects. Some minor differences are observed at resonance peaks associated with modes at frequencies 32,33 8852 and

35,36 9451Hz. From Figure 3.1, 32,33  8852 Hz are split to 32 8790 Hz and

101

5000

without gyroscopic effect (a) - - - - with gyroscopic effect 4000

3000

2000

1000

0 7000

6000 (b)

5000

4000

3000

2000

1000

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 3.5: Maximum dynamic mesh force amplitudes at the left side (a) s- pi and (b) r-pi meshes with and without gyroscopic effect.

102

32 8901Hz at m 8850 Hz (corresponding to s 11,300 rpm) when the gyroscopic effects are included. The above variation in natural frequency is marked by line p1-p1 in Figure 3.1. This change in natural frequency is reflected in Figure 3.5 with a shift in resonance frequency from 8852 Hz to 8790 Hz. Similarly shift in resonance peak at 9451 Hz can be explained by the change in the natural frequency marked by line p2-p2 in Figure 3.1.

The same analysis is repeated for the case of radially floating sun gear with a 50% stagger ( stg ). Figure 3.6 shows dynamic mesh forces for sun-planet and ring-planet meshes on left side, with and without gyroscopic effects included. It can be seen that resonance peak associated with 14  891 Hz shows considerable change in amplitude, accompanied by a slight change in resonant frequency. Including gyroscopic effects at this frequency ( m 891 Hz or s 1,137 rpm), splits one pair of these four modes to

1 881 Hz and 4 902 Hz while the other two modes remain unchanged at

2,3 891 Hz. The rest of the excited modes do not exhibit any change in natural frequency due to gyroscopic effect. As a result, the impact of the gyroscopic effects in the force response of the example gear set can be deemed secondary.

3.5 Summary

In this chapter, a certain class of gyroscopic effects is included in the linear time- invariant model of Chapter 2. Gyroscopic moments due to the resistance of a spinning gear to its change in plane of rotation are incorporated in the model. Free and forced 103

8000 without gyroscopic effect (a) - - - - with gyroscopic effect

6000

4000

2000

0 10000

(b) 8000

6000

4000

2000

0 0 200040006000800010000 (Hz)

Figure 3.6: Maximum dynamic mesh force amplitudes at the left side (a) s- pi and (b) r-pi meshes with and without gyroscopic effect for radially floating sun ( ).

104 vibration analyses are carried out for the resulting set of equations by solving complex

Eigen value problem and applying modal summation technique to the state-space form of the system equations. The example simulations indicate that only a portion of the natural modes, those exhibiting significant tilting motions, are impacted by gyroscopic moments.

Softer tilting bearing support stiffness conditions and larger polar mass moments of inertia were shown to increase the impact of gyroscopic effects on these natural frequencies. Inclusion of the gyroscopic effects were shown to influence the forced frequency response of the example gear set very little, as certain resonant peak amplitudes and frequencies are varied slightly.

References for Chapter 3:

[3.1] Meirovitch, L., 2001, Fundamentals of Vibration, McGraw-Hill Higher Education, NY.

105

CHAPTER 4

Investigation of Time-Varying Gear Mesh Stiffness Effects on Dynamics of

Double-Helical Planetary Gear Sets

4.1 Introduction

Models presented in Chapters 2 and 3 employed two related assumptions in modeling of the gear meshes. First was the assumption of constant mesh stiffness values.

It is well-established through accurate gear contact load distribution models that the overall stiffness of a gear mesh varies with roll angle (or time) in a periodic manner. The most apparent reason for such parametric fluctuation in the stiffness of a gear mesh is the fact that number of tooth pairs in contact fluctuates between two integers. A pair with a typical profile contact ratio value has one tooth pair of two tooth pairs in contact at a given time. With the stiffness of a tooth pair is relatively insensitive to the rotational position, the overall gear mesh stiffness is either formed by single tooth pair stiffness (in the zone of single-tooth contact) or two in-parallel tooth pair springs. In line with this, time dependency of a mesh stiffness of a spur gear pair was experimentally demonstrated to impact the steady-state response amplitudes of a spur gear significantly

106

[4.1]. It was also shown experimentally by the same investigators that mesh stiffness fluctuations are also responsible for other unique behavior such as parametric resonances

[4.2]. With such a experimental evidence in place, various modeling studies (e.g. refs.

[4.3-4.7]) were shown to correlate with spur gear pair measurements closely. However such time-varying stiffness effects were discounted by experimental studies on helical gear pairs including Kubur et al [4.8] showed very good match between the predictions from a three-dimensional Linear Time-Invariant (LTI) model of a helical gear system to their own helical gear pair measurements. Various three-dimensional helical gear models adapted this LTI approach (e.g. [4.9-4.11]). This was deemed reasonable since the loads in helical gear meshes are carried by a larger number of tooth pairs and contact lines are diagonal to the tooth surfaces, the overall stiffness fluctuations of helical meshes can be significantly smaller than their spur counterparts.

The second assumption used in Chapters 2 and 3 was that the gears forming the pair maintain their contact all the time, i.e. no tooth separations are allowed to take place. As any gear pair must have certain amount of backlash (clearance) for other design reasons such as lubrication and assembly, one would expect that the teeth should loose contact as soon as the dynamic gear mesh force amplitudes exceed the mean gear mesh force representing the transmitted torque. Experiments of Munro [4.12], Kubo [4.13] as well as more recent measurements by Kahraman and Blankenship [4.1-4.4, 4.14] all indicate that such tooth separations are common for spur gear pairs especially near the primary and sub-harmonic resonance peaks where dynamic gear mesh force amplitudes become large. As such, softening type forced response curves with discontinuities were reported

107 in these experiments as well as numerous Nonlinear Time-Varying (NTV) modeling studies published within the last two decades. Ma and Kahraman [4.15, 4.16] described these interactions between the parametric and nonlinear effect on a single-degree-of- freedom oscillator theoretically. Such nonlinear behavior is not evident for helical gear pairs, as the limited published studies [4.8, 4.17] indicate that the mesh stiffness fluctuations and the transmission error amplitudes are not large enough to trigger nonlinear behavior.

Such simplifying assumptions that reduce the gear dynamics model to LTI ones can be justified to a certain extent in view of published helical gear experiments, real motivation in using them often been the conveniences afforded by linear systems. The

Eigen value solutions leading to the forced response using frequency domain analysis makes larger-scale LTI system very desirable for computational purposes. The solution of the NTV model of a double-helical planetary gear set must rely on time-domain numerical integration techniques for their solutions. This chapter proposes an NTV model of a double-helical planetary gear set with periodic gear mesh stiffness conditions and tooth separations to investigate whether these effects play a tangible role on the forced response. For this, the LTI model of Chapter 3 will be modified in the next section to include mesh stiffness functions with phasing relations similar to static transmission error functions as well as subjecting them to a piecewise-linear backlash function. The

NTV model results will be compared at the end to the LTI results to assess the significance of these effects for double-helical planetary gear sets.

108

4.2 A Nonlinear Time-Varying Dynamic Model

For computational concerns, it has been a common practice to reduce the degrees of freedom of the gear dynamics model when time-varying and non-linear effects are included. The NTV models of planetary gear sets are often torsional [4.7] or two- dimensional [4.18]. Such a reduction to the LTI model of Chapters 2 and 3 is not prudent here as the motions predicted were truly three-dimensional as evident from the components of the gear mesh strain energies as well as the excited mode shapes. With this, the task in hand is to modify the 18(N  3) degree-of-freedom model of the double- helical planetary gear set (N is the number of planet branches) from Chapter 2 with the gyroscopic effects from Chapter 3 to include the time-varying gear mesh stiffnesses and tooth separations.

4.2.1 Definition of Time-Varying Gear Mesh Stiffnesses

The excitations for the LTI model consisted of the motion transmission errors applied in the form of periodic displacement functions. For this, a gear load distribution model [4.19] was used to predict the transmission error excitation at an individual external (sun-planet) and internal (ring-planet) mesh. In Section 2.2.6, these excitations were written in Fourier series form and modified to impose planet mesh phasing conditions determined by the planet position angles (  pi with p1 0 ), number of planets (N), the number of teeth on the sun and ring gears ( Zs and Zr ), and the stagger

109 between the right and left sides of the gear set ( stg ). The same methodology will be applied here to the gear mesh stiffness functions as well.

With the first sun-planet mesh on the left side as the reference mesh as in Section

2.2.6, the mesh stiffness of this mesh is predicted using the load distribution model [4.19] and written in Fourier series form as

L ()L ˆ ktksp1 ()spsplmspl k cos( lt  ). (4.1a) l1 where kˆ and  are the amplitude and phase angle of the l-th harmonic component, spl spl

()L and ksp is the mean value. Here, ktsp1 () is simultaneous with the transmission error excitation in the same mesh, given in Chapter 2 (Eq. (2.22a) as

L ()L etsp1 () eˆspl cos( lt m spl ), (4.1b) l1 where eˆ and  are the amplitude and phase angle of the l-th harmonic. In other spl spl

()L ()L words, phase angles spl and spl are such that ktsp1 () and etsp1 () at any time t represent the same exact gear mesh position.

The gear mesh stiffness at other s-pi meshes (iN[2, ]) on the left side are defined relative to the reference s-p1 mesh as

110

L ()L ˆ ktkspi ()sp k spl cos( lt  m spl l spi ), i  [2, N ], (4.2a) l1 where

Zspi , for CW planet rotation, spi  (4.2b) Zspi, for CCW planet rotation.

With  representing intentional nominal stagger of the teeth between the right stg

and left sides, stiffness functions on the right side sun-planet meshes are written as

L ()R ˆ ktkspi ()sp k spl cos( lt  m spl l spi l stg ), i  [1, N ]. (4.3) l1

The gear mesh stiffness at the r-p1 mesh on the left side is defined in relation to the reference s-p1 mesh of the left side as

L ()L ˆ ktkrp1 ()rp k rpl cos( lt  m rpl l rs ), (4.4) l1 where  is phase difference between the reference s-p1 mesh and the r-p1. Here, kˆ rs rpl and  are the amplitude and phase angle of the l-th harmonic of the mesh stiffness and rpl

krp is mean value of gear mesh stiffness for ring planet mesh. Similarly, mesh stiffness

of other r-pi meshes on the left side are given as

L ()L ˆ ktkrpi ()rp k rpl cos( lt  m rpl l rpi l rs ), i  [2, N ], (4.5a) l1 111 where

Zrpi, for CW planet rotation, rpi  (4.5b) Zrpi , for CCW planet rotation.

Finally, with the same stagger stg , gear mesh stiffness functions for ring-planet meshes on the right side of the gear set are defined as

L ()R ˆ ktkrpi ()rp k rpl cos( lt  m rpl l rpi l rs l stg ), i  [1, N ]. (4.6) l1

 In above equations, the gear mesh frequency m can be obtained using same kinematic relationship as defined in Eq. (2.24).

4.2.2. Equations of Motions for NTV Model

Equations of motion of the Linear Time Invariant (LTI) model with gyroscopic effects are modified here to take into account time-varying mesh stiffness and tooth separation nonlinearities.

With the same set of parameters, equations of motion for sun-planet i pair as defined in Eq. (2.2a-c, f and 3.4a,b) and Eq. (2.3a-c, f and 3.5a,b) are modified to include time varying mesh stiffness and tooth separation nonlinearities. The new set of equations for sun gear motion are given as

mys s() t h spi k spi ()cos t cos spi p spi () t 0, (4.7a)

112

mxss () t h spispi k ()cossin t spispi p () t 0, (4.7b)

mzss () t h spispi k ()sin t p spi () t 0, (4.7c)

ItJsys() s sxs  () thktr spispi () s sin  cos spispi pt ()  0, (4.7d)

ItJsxs() s sys  () thktr spispi () s sin  sin spispi pt ()  0, (4.7e)

Jszs()thktr spispi () s cos  ptT spi ()  s /(2 N ). (4.7f)

The corresponding equations of motion for planet i are also modified as

myp pi() t h spi k spi ()cos t cos spi p spi () t  0, (4.8a)

mxppi () t h spispi k ()cos t sin spispi p () t 0, (4.8b)

mzp pi() t h spi k spi ()sin t p spi () t 0, (4.8c)

ItJthktrp ypi() p p  xpi () spi spi () p sin  cos spi pt spi ()  0, (4.8d)

ItJp xpi() p p  ypi () thktr spi spi () p sin  sin spi pt spi ()  0, (4.8e)

Jthktrptp zpi() spi spi () p cos  spi ()  0. (4.8f)

The gear mesh stiffness term ktspi () carries superscript L or R based on whether left of right side of double-helical gear set is considered. The term h in these equations is the spi unit step function representing tooth separation as

1, ptspi ( ) 0, hspi   (4.9) 0, ptspi ( ) 0.

113

A negative/zero relative mesh displacement ptspi () 0 represents a tooth separation condition, resulting in a zero gear mesh spring force. The back collisions of the teeth are not included in this formulation based on the experimental evidence provided in Ref.

[4.1-4.4]. Equations of motion, Eq. (4.7) and (4.8) can be written in matrix form

M0qssss ()tt G0q   ()    0Mppi  q()tt  0G ppi  q  ()

11 12 KKspi spi q s()tht f sm spi f si ()  hk() t   . (4.10) spi spi 22 qf()tht () sym. Kspi pi spi spi 

In similar manner, equations of motion given by Eq. (3.7a) for a ring planet i sub- system are modified to obtain

M0qrrrr ()tt G0q   ()    0Mppi  q()tt  0G ppi  q  ()

11 12 KKrpi rpi q r()tht f rm rpi f ri ()  hk() t   , (4.11a) rpi rpi 22 qf()tht () sym. Krpi pi rpi rpi  where

1, ptrpi ( ) 0, hrpi   (4.11b) 0, ptrpi ( ) 0.

The formulation for a carrier-planet i sub-system given in Eq. (3.8a) is not influenced by the time-varying and nonlinear effects, which is repeated here for completeness purposes: 114

M0q()tt 00q () KK11 12 q () t cc    ccpicpic fcm     . 0Mppi q()tt 0G ppi q  ()22 q pi () t0    sym. Kcpi  (4.12)

As it is done for the LTI formulation, these equations of motion are applied to an entire double-helical planetary gear set configurations to obtain the overall equations of motion including time-varying gear mesh stiffnesses, tooth separation nonlinearities and gyroscopic effects as

Mq()tt + [ C () +G ] q  () t + [ Kmesh ( q ,)+ t K b ] q () t F m  F (). t (4.13a)

Here it is noted that both the stiffness matrix and damping matrix are time-varying since a proportional damping is used in the form

CK+M()tt () . (4.13b)

4.3 Solution of the NTV System Equations

A direct numerical time integration scheme is employed to solve the equations governing the NTV system. As most of the numerical time integration algorithms are written for a first order system of equations, the second order ordinary differential equations of Eq. (4.13a) are first converted to first order differential equations by writing them in state-space form as

rAr+BF+F()ttt () () [m ()], t (4.14a)

115 with the state vector r and matrices A and B defined as

q()t r()t   , (4.14b) q()t

0I  0  A()t  , B= . (4.14c,d) 11  1 MKq(,)tt M[G+C] () M 

Here K(,)q tt Kmesh (,)+q K b .

This system of 36(N  3) first order NTV ordinary differential equations is solved using Gear’s BDF (Backward Differentiation Formulae) method [4.20]. Due to large scale of the problem (Eigen values for the corresponding LTI system ranges from 0 to

1012 ), large differences in stiffness properties of the system and large number of degrees of freedom, equations of motion behave stiff, requiring a stiff solver as the one used here. The numerical integration must be carried out for extended periods to pass the transients such that the steady-state response can be obtained. As in any dynamic system, larger damping levels and more appropriate selection of initial conditions r(0) should shorten the transient region of the response. It is also expected that the transients last longer in the vicinity of the resonance peaks than off resonance regions. While not much can be done about damping and proximity to the resonance peaks, the initial conditions at a given gear mesh frequency (speed) increment were chosen as the final steady-state solution from the previous mesh frequency increment to minimize the simulations required to avoid transient solutions.

116

The state vector r()t consists of displacement vector q()t and the velocity vector q()t . With q()t known, relative gear mesh displacements are computed according to Eq.

(2.4) and (2.9), from which dynamic mesh forces at each of the 4N gear meshes can be obtained. The dynamic factors as defined by Eq. (2.31) are valid here as well.

4.4 Numerical Results

The NTV model formulation employs two separate excitations at a gear mesh j, a static transmission error excitation et() and a parametric gear mesh stiffness excitation j kt(). The static transmission error et() has a load-depended component et() that j j jd represents the tooth deflections and a kinematic component etjk () that represents the intentional tooth modifications and manufacturing errors, both of which cause the tooth surfaces to deviate from a perfect involute. With this, etjjdjk() e () t e () t.

In Chapters 2 and 3 where ktj () kjm  constant, both components of the etj () were included by considering the loaded static transmission error (LSTE) as etj (). This was proposed by Ozguven and Houser [4.21] as a good approximation for spur gears as well. Focusing on helical gear mesh modeling, Blankenship and Singh [4.22] suggested that using LSTE along with a time-varying gear mesh stiffness might not be correct as the same tooth deflections contribute to both et() and kt(). Per Tamminana et al [4.23] j j four versions of gear mesh models for spur gear pair: (i) ktj () and etjjdjk() e () t e () t

, (ii) ktj () and etjjk() e () t, and (iii) ktj () and etj () 0. Through comparison to 117

NTV spur gear responses and predictions of a 2D deformable-body dynamics model,

Tamminana et al [4.23] recommended that second option with et() e () tshould be jjk used for spur gear pairs having tooth modifications while the third option with etj () 0 is the best in case of unmodified (perfect involute) gears.

No such extensive study on the gear mesh modeling exists for helical gears for two reasons. One is that only a very small fraction of published gear dynamics data is available for helical gears. In addition, there is no 3D deformable-body model available to simulate helical gears under dynamic conditions. For this reason, the NTV model proposed in Section 4.2 will be exercised with four variations of excitation parameters:

Model I: LTI system with ktj () kjm  constant and etjjdjk() e () t e () t ,

Model II: NTV system with ktj () and etjjdjk() e () t e () t,

Model III: NTV system with ktj () and etjjk() e () t, and

Model IV: NTV system with ktj () and etj () 0.

4.4.1 Verification and Analysis of Time-domain Solutions

The numerical time integration scheme employed to solve Eq. (4.14a) is first verified by comparing its steady-state solutions for the limiting LTI case to those from the frequency-domain modal summation solution from Chapter 2. Figure 4.1 compares the maximum dynamic gear mesh force amplitudes predicted for the example system defined in Table 2.1 for the same proportional damping values of (1.35(10)6 s, 118

 50 s-1) with sun torque of 2000 Nm. Here numerical integration solutions were carried out for 500 gear mesh periods (tTmax  500 mesh whereTmesh2 m ) to pass the transient region such that steady-state response can be obtained for this comparison.

As shown in Figure 4.1(a) and (b), respectively, both dynamic sun-planet and ring-planet mesh forces predicted by the two methods match perfectly for the limiting LTI case indicating that the state-space form of the equations of motion as well as the numerical integration scheme works accurately.

4.4.2 Example System Analyses

In this section, steady-state responses obtained by all three NTV gear mesh models specified in Section 4.4 will be compared to each other as well as that of the corresponding LTI system. With the transmission error and mesh stiffness excitation parameters defined in Table 4.1 for sun torque of 2000 Nm (Ts  2000 Nm), Figure 4.2 provides such a comparison for the same system used in Figure 4.1. Here the amplitude of the response is the root-mean-square values of the dynamic gear mesh forces and is defined as

L ˆ 2 ()Fpi rms Fsr pil , ,. (4.15) l1

ˆ where Fpil is the l-th harmonic of the steady-state gear mesh force. It is seen in this figure that the r.m.s. dynamic amplitudes follow the same trends exciting similar set of

119

2000

(a) - - - -

1600

1200

(N) 800

400

0 3000

(b)

2400

1800 (N)

1200

600

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 4.1: Comparison of modal summation and direct numerical integration solutions for the LTI case. Maximum dynamic mesh force amplitudes on the left side (a) s-pi and (b) r-pi meshes.

120

Table 4.1: Harmonic amplitudes and phase angles of the transmission error and mesh stiffness excitations of the example gear set of Table 2.1 [4.19] for different torque levels.

Ts  2000 Nm

eˆ kˆ eˆ kˆ     l spl spl rpl rpl spl spl rpl rpl ( µm) ( N/µm) (µm) (N/µm) (rad) (rad) (rad) (rad) 1 0.405 36.55 0.483 36.41 -0.769 2.429 -0.709 2.501 2 0.088 5.88 0.135 7.89 -1.299 1.981 -1.543 1.674 3 0.021 1.09 0.005 0.01 0.604 -2.291 0.927 -1.899

Ts 1000 Nm

eˆ kˆ eˆ kˆ     l spl spl rpl rpl spl spl rpl rpl ( µm) ( N/µm) (µm) (N/µm) (rad) (rad) (rad) (rad) 1 0.068 32.08 0.09 31.74 -0.75 2.49 -0.75 2.56 2 0.025 5.87 0.04 7.62 -1.19 2.09 -1.19 1.79 3 0.014 1.15 0.01 0.26 0.760 -2.06 0.76 -1.76

Ts  667 Nm

eˆ kˆ eˆ kˆ     l spl spl rpl rpl spl spl rpl rpl ( µm) ( N/µm) (µm) (N/µm) (rad) (rad) (rad) (rad) 1 0.048 24.73 0.032 24.44 2.36 2.46 2.59 2.58 2 0.005 7.01 0.008 8.46 -0.52 2.00 -1.48 1.85 3 0.011 1.22 0.011 0.25 0.85 -2.05 0.76 -1.66

121 modes at same resonance frequencies while the amplitudes of certain resonance peaks differ with the gear mesh model used. In Figure 4.2, the peak at

11(8398)  4199Hz is excited for mesh models I and II suggesting that m 22 transmission error component of the excitations excite this mode. Resonance peak at

8852 m  Hz are excited somewhat equally regardless of the gear mesh model used, while the peak at 11(11047)  5523 Hz is excited in case of models II, m 22

III and IV indicating that mesh stiffness is responsible for exciting this mode. Meanwhile the peak at m  9451Hz is seen for Models II, III and IV indicating again that mesh stiffness variation is responsible for exciting this mode.

Some of the time histories of the steady state response are presented next to demonstrate their shape as well as harmonic content. In relation to Figure 4.2, Figure

4.3(a) shows dynamic gear mesh force time history for the left side s-p1 mesh at

8852 m Hz, which represents a resonance peak in Figure 4.2. As evident from the

Fast Fourier spectrum of the same data shown in Figure 4.3(b), the response at

m 5523 Hz is dominated by the second gear mesh harmonic, indicating that this mode is excited by the second harmonic terms of excitations

The root-mean-square representation of the steady-state response used in Figure 4.2 was not used in earlier chapters. Instead, maximum dynamic response amplitude defined in Chapter 2 was used. Using the same maximum dynamic mesh force amplitudes, Figure

4.2 is reproduced in Figure 4.5. Comparison of these two figures indicate that the

122

2500

(a)

2000

1500

(N) 1000

500

0 3000

(b) 2400

1800

(N) 1200

600

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 4.2: Root-mean-square values of dynamic mesh forces at the left side (a) s-pi and (b) r-pi meshes for different excitation models ().

123

9000 (a)

8000

7000

(N) 6000

5000

4000 0 1 2 3 4 Mesh cycles

2500 (b)

2000

1500

(N) 1000

500

0 0 1 2 3 4 5 Mesh order

Figure 4.3: Steady-state dynamic mesh force at an s-pi mesh on the left side obtained by using Model II. (a) Time history at Hz and (b) the corresponding frequency spectrum.

124

7500 (a)

7000

6500

(N) 6000

5500

5000 0 1 2 3 4 Mesh cycles 1000 (b)

800

600

(N) 400

200

0 0 1 2 3 4 5 Mesh order

Figure 4.4: Steady-state dynamic mesh force at an s-pi mesh on the left side obtained by using Model II. (a) Time history at Hz and (b) the corresponding frequency spectrum.

125

2500 (a)

2000

1500

(N) 1000

500

0 3000

(b) 2400

1800

(N) 1200

600

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 4.5: Maximum dynamic mesh force amplitudes at the left side (a) s-pi and (b) r-pi meshes with different excitation models ( ).

126 response curves are very similar for both cases such that r.m.s. or maximum gear mesh force amplitudes can be used as the metric for the forced response plots.

Figure 4.6 shows maximum amplitudes of dynamic mesh forces for same gear set with 50% stagger ( stg ) between the right and left sides of the gear set. All for gear mesh models exhibit resonance peaks at   2999Hz, with models II to IV causing higher resonance peaks as shown in Figure 4.6(b).

Next, an In-phase gear set analyzed in Chapter 2 is revisited here to examine the influence of different gear mesh models on the system response. Figure 4.7 shows the maximum dynamic mesh forces for this in-phase system for all four types of gear mesh models. As observed earlier, resonance frequencies remain same with amplitude of response varying considerably across models I to IV.

As the final analysis, influence of torque levels on the dynamic mesh forces is shown in Figure 4.8 using Model II as an example. For the example system of Table 2.1, the excitation parameters change with torque as shown in Table 4.1 for three sun gear torque values of Ts  2000 , 1000 and 667 Nm. It can be seen from Figure 4.8 that the shapes of the resonance curves remain same but amplitudes increase with an increase in mean torque levels. This is partly because the excitation parameters vary with torque.

More importantly, the increase in mean torque increases the influence of time-varying stiffness excitations, which is in agreement with References [4.1, 4.2, 4.24].

127

10000 (a)

8000

6000

(N) 4000

2000

0 8000

(b)

6000

4000 (N)

2000

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 4.6: Maximum dynamic mesh force amplitudes at the left side (a) s-pi and (b) r-pi meshes for different excitation models ( ).

128

5000

(a)

4000

3000

(N) 2000

1000

0 5000 (b)

4000

3000

(N) 2000

1000

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 4.7: Maximum dynamic mesh force amplitudes at the left side (a) s-pi and (b) r-pi meshes of an in-phase system for different gear mesh models ( ).

129

2500 (a)

2000

1500

(N) 1000

500

0 2500

(b) 2000

1500

(N) 1000

500

0 0 2000 4000 6000 8000 10000 (Hz)

Figure 4.8: Maximum dynamic mesh force amplitudes at the left side of (a) s-pi and (b) r-pi mesh at various input torque levels. Model II was used here.

130

4.5 Summary

This chapter is focused on potential effects of time-varying gear mesh stiffness and nonlinearities on the steady-state response of a double-helical planetary gear set. The model preserves generic nature of the existing model such that any number of planets, any phasing condition, spacing configurations and any support conditions can still be analyzed. Direct numerical integration scheme was employed to solve the resulting set of nonlinear time varying equations of motion in time domain. Based on past studies on discrete dynamic modeling of helical gear pairs, four gear mesh model variations (the first one being the LTI system of Chapter 2) were defined and the model was exercised with all mesh models. Sizable differences were observed in forced response curves with inclusion of time-varying stiffnesses, characterized by new resonance peaks and overall increased response amplitudes. In the absence of detailed experimental data, it was not possible to discriminate for or against any of these gear mesh models. However their closeness to the LTI solution suggests that the LTI system approximation might be sufficient for practical design purposes, which is in agreement with the conclusions of

Kubur et al [4.8] for regular helical gears.

While the variations of the model used in this chapter all implemented tooth separation functions defined by Eq. (4.9) and (4.11b) to allow contact loss in cases the relative gear mesh displacements pspi ()t and prpi ()t reach zero. None of the simulations presented however show any signs of nonlinear behavior. There are no jump discontinuities and peaks exhibit symmetric, linear behavior. This confirms the

131 conclusions of Kubur et al [4.8] for regular helical systems. In view of these simulations, the model can be reduced to a linear time-varying one by setting hhspi rpi 1 in Eq.

(4.9) and (4.11b).

References for Chapter 4:

[4.1] Kahraman, A. and Blankenship, G. W., 1999, “Effect of Involute Contact Ratio on Spur Gear Dynamics,” ASME Journal of Mechanical Design, 121, pp. 112- 118.

[4.2] Kahraman, A. and Blankenship, G. W., 1997, “Experiments on Nonlinear Dynamic Behavior of an Oscillator with Clearance and Time-varying Parameters,” ASME Journal of Applied Mechanics, 64, pp. 217-226.

[4.3] Kahraman, A. and Blankenship, G. W., 1996, “Interactions between Commensurate Parametric and Forcing Excitations in a Systems with Clearance,” Journal of Sound and Vibration, 194, pp. 317-336.

[4.4] Blankenship, G. W. and Kahraman, A., 1995, “Steady State Forced Response of a Mechanical Oscillator with Combined Parametric Excitation and Clearance Type Nonlinearity,” Journal of Sound and Vibration, 185(5), pp. 743-765.

[4.5] Al-Shyyab, A. and Kahraman, A., 2005, “Non-Linear Dynamic Analysis of a Multi-Mesh Gear Train using Multi-term Harmonic Balance Method: Period-one Motions,” Journal of Sound and Vibration, 284, pp. 151-172.

[4.6] Al-Shyyab, A. and Kahraman, A., 2005, “Non-Linear Dynamic Analysis of a Multi-Mesh Gear Train using Multi-term Harmonic Balance Method: Subharmonic motions,” Journal of Sound and Vibration, 279, pp. 417-451.

[4.7] Al-shyyab, A. and Kahraman, A., 2007, “A Non-linear Dynamic Model for Planetary Gear Sets,” Proc. ImechE, Part K: Journal of Multi-Body Dynamics, 221, pp. 567-576.

132

[4.8] Kubur, M., Kahraman, A., Zini, D., and Kienzle, K., 2004, “Dynamic Analysis of a Multi-shaft Helical Gear Transmission by Finite Elements: Model and Experiment,” ASME Journal of Vibrations and Acoustics, 126, pp. 398-406.

[4.9] Kahraman, A., 1994, “Dynamic Analysis of a Multi-mesh Helical Gear Train,” ASME Journal of Mechanical Design, 116, pp. 706-712.

[4.10] Kahraman, A., 1994, “Planetary Gear Train Dynamics,” ASME Journal of Mechanical Design, 116, pp. 713-720.

[4.11] Kahraman, A., 1993, “Effect of Axial Vibrations on the Dynamics of a Helical Gear Pair,” ASME Journal of Vibration and Acoustics, 115, pp. 33-39.

[4.12] Munro, R. G., 1962, “The Dynamic Behaviors of Spur Gears,” PhD Dissertation, Cambridge University, UK.

[4.13] Kubo, A., Yamada, K., Aida, T., and Sato, S., 1972, “Research on Ultra High Speed Gear Devices (reports 1-3),” Transaction of Japan Society of Mechanical Engineer, 38, pp. 2692-2715.

[4.14] Kahraman, A. and Blankenship, G. W., 1999, “Effect of Involute Tip Relief on Dynamic Response of Spur Gear Pairs,” ASME Journal of Mechanical Design, 121, pp. 313-315.

[4.15] Ma, Q. and Kahraman, A., 2006, “Sub harmonic Motions of a Mechanical Oscillator with Periodically Time-varying, Piecewise Non-linear Stiffness,” Journal of Sound and Vibration, 294(3), pp. 924-636.

[4.16] Ma, Q. and Kahraman, A., 2005, “Period-one motions of a Mechanical Oscillator with Periodically time-varying, Piecewise Non-linear Stiffness,” Journal of Sound and Vibration, 284, pp. 893-914.

[4.17] Umezawa, K., Ajima, T., and Houjoph, H., 1986, “Vibration of Three Axis Geared System,” Transaction of Japan Society of Mechanical Engineer, 29, pp. 950-957.

133

[4.18] Kahraman, A., 1994, “Load Sharing Characteristics of Planetary Transmissions,” Mechanisms and Machine Theory, 29, pp. 1151-1165.

[4.19] LDP Gear Load Distribution Program, 2011, Gear and Power Transmission Research Laboratory, The Ohio State University, USA.

[4.20] Gear, C. W., 1971, “The Simultaneous Numerical Solution of Differential- Algebraic Equations,” IEEE Trans. Circuit Theory, TC-(18), pp. 89-95.

[4.21] Ozguven, H., and Houser, D., 1988, “Dynamic Analysis of High Speed Gears by Using Loaded Static Transmission Error,” Journal of Sound and Vibration, 125(1), pp. 71-83.

[4.22] Blankenship, G. W., and Singh, R., 1995, “A New Gear Mesh Interface Dynamic Model to Predict Multi-Dimensional Force Coupling and Excitation,” Mechanism and Machine Theory, 30(1), pp. 43-57.

[4.23] Tamminnana, V. K, Kahraman A. and Vijayakar, S., 2007, “On the Relationship between the Dynamic Factors and Dynamic Transmission Error of Spur Gear Pairs,” ASME Journal of Mechanical Design, 129, pp. 75-84.

[4.24] Al-Shyyab, A., 2003, “Nonlinear Analysis of Multi Mesh Gear Train Using Multi-Term Harmonic Balance Method,” PhD Dissertation, The University of Toledo, USA.

134

CHAPTER 5

Conclusions

5.1 Summary

This research is focused on investigating the dynamic response of double-helical planetary sets through a development of an analytical model. The proposed model used a three-dimensional discrete-parameter formulation with all gear mesh, bearing and support structure compliances included. The model was presented in three levels of complexity, starting with linear time-invariant (LTI) model, then the same LTI formulation with gyroscopic effects included, and finally a nonlinear time-varying (NTV) version with parametrically time-varying gear mesh stiffnesses and nonlinear tooth separation effects included.

As the first step, a generic three-dimensional linear (no tooth separations), time- invariant (constant gear mesh stiffnesses) model was proposed to simulate dynamic behavior of any N-planet double-helical planetary gear system. The model allowed any planet phasing conditions dictated by the number of planets, number of gear teeth and planet position angles. In addition, gear mesh phasing conditions associated with any 135 stagger angle between the right and left sides of the gear set were also accounted for in the model. As excited by the loaded gear mesh motion transmission errors applied at the gear meshes as displacement excitations, forced response of the gear set was computed by using the modal summation technique, with the natural modes found from the corresponding real Eigen value solution for the undamped system. As the least known set of parameters, damping at the gear mesh and bearing interfaces were represented by a proportional damping formulation. Strain energies of the mode shapes were computed to identify the modes excitable by the transmission error excitations. Parametric studies on an example gear set showed significant influences of planet phasing, stagger conditions, gear and carrier support conditions as well as the number of planets on the steady-state forced response.

High-speed double-helical planetary gear set applications requires an investigation of the influence of gyroscopic effects on the system response. In the second step, LTI model was modified to include a class of gyroscopic effects due to vibratory skew of spinning gears, with the carrier being stationary. The corresponding complex Eigen value solutions were examined to quantify the influence of rotational speed of the gear set through gyroscopic effects on the natural frequencies and the mode shapes. A complex modal summation formulation was employed to compute the forced response with gyroscopic effects. Influence of gyroscopic moments on natural frequencies was found to be modest within typical aerospace speed ranges, with only a sub-set of modes exhibiting dominant tilting motions impacted by the gyroscopic effect. In the forced

136 response curves, effect of gyroscopic moments was limited to modest changes in amplitudes and frequencies of certain resonance peaks.

As the final step in the model development, mesh stiffness fluctuations due to change in number of tooth pairs in contact were included in the second version of the model. The stiffness fluctuations serve as internal parametric excitation, applied at the gear meshes along with the transmission error excitations with the same mesh-to-mesh phasing relations. Tooth separation nonlinearities were also included here to arrive at a set of NTV equations of motion in state-space form, which were solved by using direct numerical integration. Based on past studies on discrete dynamic modeling of helical gear pairs, three variations of transmission error excitations (zero, unloaded and loaded) were considered with the time-varying gear mesh stiffnesses. Sizable differences were noted between the forced response curves for time-varying and time-invariant systems characterized by additional resonance peaks and overall increases in response amplitudes while no signs of nonlinear behavior were evident.

5.2 Conclusions and Contributions

This research provided a mathematical tool to predict dynamic behavior of a double-helical planetary gear system. Based on the results of the simulations of example systems using this tool, a number of conclusions can be made in regards to the dynamic behavior of double-helical planetary gear sets:

137

 Natural modes exhibit shapes that can allow a classification based on the

planet gear displacements relative to the central members. Certain modes can

be characterized as axi-sysmmetric with same planet motions relative to sun

gear while the others are not.

 Numerous natural modes are predicted to have significant tilting and/or axial

motions, indicating that a three-dimensional model must be used in simulating

double-helical planetary gear sets.

 The modes excited by the excitation mechanisms devised, are shown to have

phasing conditions that match the imposed phasing of the excitations. The

modes with large gear mesh modal strain energies are the ones excited by the

gear mesh excitations.

 Both the right-to-left stagger angle and the planet-to-planet phasing conditions

are shown to change the forced response significantly, in the process exciting

different sets of natural modes. This indicates that an accurate dynamic model

of a double-helical planetary gear set must use the actual excitation phasing

conditions defined by the design parameters such as stagger angle, number of

planets, numbers of gear teeth and planet position angles.

 Only few of the natural modes exhibiting significant tilting motions are

impacted by gyroscopic moments while most of these modes are not the ones

excited by the gear mesh excitations. As such, influence of gyroscopic

138

moments on the forced vibration response was found to be secondary for the

systems with non-rotating carriers.

 Different forms of the transmission error excitations used in the time-varying

model showed differences in the resultant force response curves, highlighting

the need for experimental data to determine the accurate form of transmission

errors used in time-varying models.

 In agreement with the previous experimental observations from single-helical

gear sets, the double-helical planetary gear sets are predicted to exhibit linear

behavior. Accordingly, backlash-induced nonlinearities can be neglected in

modeling double-helical gear systems.

This proposed model represents the first generalized analytical tool to study the dynamic characteristics of double-helical planetary gear system. Given its generality, the model is applicable to systems with any number of planets, any support condition and any phasing configurations. Model can be useful to study the influence of key design parameters like stagger, phasing condition, number of planets on the dynamic response of the system within ranges of speed and torque of operation. While found to be not vitally influential for the example systems analyzed, the model provides the means for investigating gyroscopic effects as well as time-varying effects and nonlinearities of any double-helical planetary system, filling the void in these aspects of aerospace gearing as well.

139

5.3 Recommendations for Future Work

Assumptions made in development of this discrete analytical model can be removed to investigate their influence on the dynamic characteristics of the double helical planetary gear system. Some of the recommendations which can enhance the work presented in this study are

 The model can be complemented by incorporating different types of

manufacturing and assembly errors (gear run-out and tooth indexing errors,

planet tooth thickness errors, carrier eccentricity, planet pin hole position

errors and ring gear roundness error) to study their influence on the dynamic

characteristics of the gear system.

 Flexibility of ring gear can also be taken into account by modeling hybrid

finite element and discrete model by using techniques like model

condensation or employing complete deformable body model.

 Gyroscopic effect due to rotating carrier can also be included in the

formulation by using rotating frame of reference.

 Analytical model presented in this study must be validated through

comparisons to experiments. For this, a new experimental set-up must be

developed to execute tightly-controlled high-speed planetary gear experiments

to generate data suitable for the validation of the model.

140

BIBLIOGRAPHY

Abousleiman, V. and Velex, P., 2006, “A Hybrid 3D Finite Element/Lumped Parameter Model for Quasi-Static and Dynamic Analyses of Planetary/Epicyclic Gear Sets,” Mechanism and Machine Theory, 41, pp. 725-748.

Abousleiman, V., Velex, P., and Becquerelle, S., 2007, “Modeling of Spur and Helical Gear Planetary Drives with Flexible Ring Gears Planet Carriers,” ASME Journal of Mechanical Design, 129, pp. 95-106.

Ajmi, M. and Velex, P., 2001, “A Model for Simulating the Quasi-Static and Dynamic Behavior of Double Helical Gears,” The JSME International Conference on Motion and Power Transmission, MPT-2001, pp. 132-137.

Al-Shyyab, A., 2003, “Nonlinear Analysis of Multi Mesh Gear Train Using Multi-Term Harmonic balance Method,” PhD Dissertation, The University of Toledo, USA.

Al-Shyyab, A. and Kahraman, A., 2005, “Non-Linear Dynamic Analysis of a Multi-Mesh Gear Train using Multi-term Harmonic Balance Method: Period-one Motions,” Journal of Sound and Vibration, 284, pp. 151-172.

Al-Shyyab, A. and Kahraman, A., 2005, “Non-Linear Dynamic Analysis of a Multi-Mesh Gear Train using Multi-term Harmonic Balance Method: Subharmonic motions,” Journal of Sound and Vibration, 279, pp. 417-451.

Al-Shyyab, A. and Kahraman, A., 2007, “A Non-linear Dynamic Model for Planetary Gear Sets,” Proc. ImechE, Part K: Journal of Multi-Body Dynamics, 221, pp. 567-576.

141

Al-Shyyab, A. and Kahraman, A., 2007, “A Nonlinear Dynamic Model for Planetary Gear Sets,” Proceedings of the Institution of Mechanical Engineers Part K: Journal of Multi-Body Dynamics, 221, pp. 567-576.

Anderson, N., Nightingale, L., and Wagner, A., 1989, “Design and Test of Turbofan Gear System,” Journal of Propulsion, 5(1), pp. 95-102.

August, R. and Kasuba, R., 1986, “Torsional Vibration and Dynamic Loads in a Basic Planetary Gear System,” ASME Journal of Vibration, Acoustics, Stress, and Reliability in Design, 108, pp. 348-353.

Blankenship, G. W. and Kahraman, A., 1995, “Steady State Forced Response of a Mechanical Oscillator with Combined Parametric Excitation and Clearance Type Nonlinearity,” Journal of Sound and Vibration, 185(5), pp. 743-765.

Blankenship, G. W. and Singh, R., 1995, “A New Gear Mesh Interface Dynamic Model to Predict Multi-Dimensional Force Coupling and Excitation,” Mechanism and Machine Theory, 30(1), pp. 43-57.

Blunt, D. and Keller, J., 2006, “Detection of Fatigue Crack in a UH-60A Planet Gear Carrier using Vibration Analysis,” Mechanical Systems and Signal Processing, 20, pp. 2095-2111.

Bodas, A. and Kahraman, A., 2004, “Influence of Carrier and Gear Manufacturing Errors on the Static Load Sharing Behavior of Planetary Gear Sets,” JSME International Journal, 47, pp. 908-915.

Botman, M., 1976, “Epicyclic Gear Vibrations,” Journal of Engineering for Industry, 97, pp. 811-815.

Chaari, F., Fakhfakh, T., Hbaieb, R., Louati, J., and Hadder, M., 2006, “Influence of Manufacturing Errors on the Dynamic Behavior of Planetary Gears,” International Journal of Advance manufacturing Technology, 27, pp. 738-746.

Clapper, M. and Houser, D., 1993, “Prediction of Fully Reversed Stresses at the Base of the Root in Spur and Double Helical Gears in a Split Torque Helicopter Transmission,” Proceedings of American Helicopter Society Rotor Wing Specialists Meeting, Williamsburg, VA. 142

Cunliffe, F., Smith, J., and Welbourn, D., 1974, “Dynamic Tooth Loads in Epicyclic Gears,” ASME Journal of Engineering for Industry, 95, pp. 578-584.

Gear, C. W., 1971, “The Simultaneous Numerical Solution of Differential-Algebraic Equations,” IEEE Trans. Circuit Theory, TC-(18), pp. 89-95.

Hbaieb, R., Chaari, F., Fakhfakh, T., Hadder, M., 2006, “Dynamic Stability of a Planetary Gear Train under the Influence of Variable Meshing Stiffnesses,” Proceedings of IMechE. Part D: Journal of Automobile Engineering, 229(D12), I711-I725.

Hidaka, T. and Terauchi, Y., 1976, “Dynamic behavior Planetary Gear, 1st Report: Load Distribution in Planetary Gear,” Bulletin of JSME, 19, pp. 690-698.

Hidaka, T., Terauchi, Y., and Dohi, K., 1979, “On the Relation between the Run-Out Errors and the Motion of the Center of Sun Gear in Stoeckicht,” Bulletin of JSME, 22(167), pp. 748-754.

Hidaka, T., Terauchi, Y., and Fujii, M., 1980, “Analysis of Dynamic Tooth Load on Planetary Gear,” Bulletin of JSME, 23, pp. 315-323.

Inalpolat, M. and Kahraman, A., 2008, “Dynamic Modeling of Planetary Gears of Automatic Transmissions,” Proceedings of Institution of Mechanical Engineers Part K: Journal of Multi-body Dynamics, 222, pp. 229-242.

Inalpolat, M. and Kahraman, A., 2009, “A Theoretical and Experimental Investigation of Modulation Sidebands of Planetary Gear Sets,” Journal of Sound and Vibration, 323, pp. 677-696.

Inalpolat, M. and Kahraman, A., 2010, “A Dynamic Model to Predict Modulation Sidebands of a Planetary Gear Set having Manufacturing Errors,” Journal of Sound and Vibration, 329, pp. 371-393.

Jauregui, J. and Gonzalez, O., 1999, “Modeling Axial Vibrations in Herringbone Gears,” Proceedings of ASME Design Engineering technical conference, Nevada, DETC99/VIB- 8109.

143

Kahraman, A., 1993, “Effect of Axial Vibrations on the Dynamics of a Helical Gear Pair.” ASME Journal of Vibration and Acoustics, 115, pp. 33-39.

Kahraman, A., 1994, “Natural Modes of Planetary Gear Trains,” Journal of Sound and Vibration, 173(1), pp. 125-130.

Kahraman, A., 1994, “Dynamic Analysis of a Multi-mesh Helical Gear Train.” ASME Journal of Mechanical Design, 116, pp. 706-712.

Kahraman, A., 1994, “Load Sharing Characteristics of Planetary Transmission,” Mechanism and Machine Theory, 29(8), pp. 1151-1165.

Kahraman, A., 1994, “Planetary Gear Train Dynamics,” ASME Journal of Mechanical Design, 116, pp. 713-720.

Kahraman, A., 2001, “Free Torsional Vibration Characteristics of Compound Planetary Gear Sets,” Mechanism and Machine Theory, 36, pp. 953-971.

Kahraman, A., and Blankenship, G., 1994, “Planet Mesh Phasing in Epicyclic Gear Sets,” Proceedings of ASME Power Transmission and Gearing Conference, San Diego.

Kahraman, A. and Blankenship, G. W., 1996, “Interactions between Commensurate Parametric and Forcing Excitations in a Systems with Clearance,” Journal of Sound and Vibration, 194, pp. 317-336.

Kahraman, A. and Blankenship, G. W., 1997, “Experiments on Nonlinear Dynamic Behavior of an Oscillator with Clearance and Time-varying Parameters,” ASME Journal of Applied Mechanics, 64, pp. 217-226.

Kahraman, A. and Blankenship, G. W., 1999, “Effect of Involute Contact Ratio on Spur Gear Dynamics,” ASME Journal of Mechanical Design, 121, pp.112-118.

Kahraman, A. and Blankenship, G. W., 1999, “Effect of Involute Tip Relief on Dynamic Response of Spur Gear Pairs,” ASME Journal of Mechanical Design, 121, pp. 313-315.

144

Kahraman, A., Kharazi, A., and Umrani, M., 2003, “A Deformable Body Dynamic Analysis of Planetary Gears with Thin Rims,” Journal of Sound and Vibration, 262, pp. 752-768.

Kang, M. and Kahraman, A., 2012, “Measurement of Vibratory Motions of Gears Supported by Compliant Shafts,” Mechanical Systems and Signal Processing, 29, pp. 391-403.

Kubo, A., Yamada, K., Aida, T., and Sato, S., 1972, “Research on Ultra High Speed Gear Devices (reports 1-3),” Transaction of Japan Society of Mechanical Engineer, 38, pp. 2692-2715.

Kubur, M., Kahraman, A., Zini, D., and Kienzle, K., 2004, “Dynamic Analysis of a Multi-Shaft Helical Gear Transmission by Finite Elements: Model and Experiment,” ASME Journal of Vibration and Acoustics, 126, pp. 398-406.

LDP Gear Load Distribution Program, 2011, Gear and Power Transmission Research Laboratory, The Ohio State University, USA.

Ligata, H. and Kahraman, A., 2008, “An Experimental Study of the Influence of Manufacturing Errors on the Planetary Gear Stresses and Planet Load Sharing,” ASME Journal of Mechanical Design, 130, 041701-1 – 041701-9.

Ligata, H., Kahraman, A., and Singh, A., 2009, “Closed-form Planet Load Sharing Formulae for Planetary Gear Sets using Translational Analogy,” Journal of Mechanical Design, 131, 021007-1 to 021007-7.

Lin, J., 2000, “Analytical Investigation of Planetary Gear Dynamics,” PhD Thesis, The Ohio State University, USA.

Ma, Q. and Kahraman, A., 2005, “Period-One Motions of a Mechanical Oscillator with Periodically Time-Varying, Piecewise Non-linear Stiffness,” Journal of Sound and Vibration, 284, pp. 893-914.

Ma, Q. and Kahraman, A., 2006, “Sub harmonic Motions of a Mechanical Oscillator with Periodically Time-varying, Piecewise Non-linear Stiffness,” Journal of Sound and Vibration, 294(3) pp. 924-636.

145

McFadden, P. and Smith, J., 1985, “An Explanation for the Asymmetry of the Modulation Sidebands about the Tooth Meshing Frequency in Epicyclic Gear Vibration,” Proceedings of Institution of Mechanical Engineers, 199(C1), pp. 65-70.

Meirovitch, L., 2001, Fundamentals of Vibration, McGraw-Hill Higher Education, NY.

Mraz, S., 2009, “Gearing up for Geared Turbofan,” Machine Design by Engineers for Engineers”. (http://machinedesign.com/article/gearing-up-for-geared-turbofans-0202)

Munro, R. G., 1962, “The Dynamic Behaviors of Spur Gears,” PhD Dissertation, Cambridge University, UK.

Ozguven, H. and Houser, D., 1988, “Dynamic Analysis of High Speed Gears by Using Loaded Static Transmission Error,” Journal of Sound and Vibration, 125(1), pp. 71-83.

Parker, R., Agashe, V., and Vijayakar, S., 2000, “Dynamic Response of a Planetary Gear System using a Finite Element/Contact Mechanics Model,” ASME Journal of Mechanical Design, 122, pp. 304-310.

Parker, R. and Lin, J., 2004, “Mesh Phasing Relationships in Planetary and Epicyclic Gears,” ASME Journal of Mechanical Design, 126, pp. 365-370.

Platt, R. and Leopold, R., 1996, “A Study on Helical Gear Planetary Phasing Effects on Transmission Noise,” VDI Berichte, 1230, pp.793-807.

Saada, A. and Velex, P., 1995, “An Extended Model for the Analysis of the Dynamic Behavior of Planetary Trains,” ASME Journal of Mechanical Design, 117, pp. 241-247.

Seager, D., 1975, “Conditions for the Neutralization of Excitation by the Teeth in Epicyclic Gearing,” Journal Mechanical Engineering Science, 17(5), pp. 293-298.

Tamminnana, V. K, Kahraman A., and Vijayakar, S. “On the Relationship Between the Dynamic Factors and Dynamic Transmission Error of Spur Gear Pairs” ASME Journal of Mechanical Design, 129, pp. 75-84.

146

Thomas, J., 1991, “A Procedure for Predicting the Load Distribution and Transmission Error Characteristics of Double Helical Gears,” MS Thesis, The Ohio State University.

Toda, A. and Botman, M., 1979, “Planet Indexing in Planetary Gears for Minimum Vibrations,” ASME paper, 79-DET-73.

Umezawa, K., Ajima, T., and Houjoph, H., 1986, “Vibration of Three Axis Geared System,” Transaction of Japan Society of Mechanical Engineer, 29, pp. 950-957.

Velex, P. and Flamand, L., 1996, “Dynamic Response of Planetary Trains to Mesh Parametric Excitations,” ASME Journal of Mechanical Design, 118, pp. 7-14.

Vijayakar, S., 1991, “A Combined Surface Integral and Finite Element Solution for a Three-Dimensional Contact Problem,” International Journal for Numerical Methods in Engineering, 31, pp. 525-545.

Wang, C., Fang, Z., and Jia, H., 2010, “Investigation of Design Modification for Double Helical Gears Reducing Vibration and Noise,” Journal of Marine Science and Applications, 9, pp. 81-86.

Yang, J. and Dai, L., 2008, “Survey of Dynamics of Planetary Gear Trains,” International Journal of Materials and Structural Integrity, 1, pp. 302-322.

Yuksel, C. and Kahraman, A., 2004, “Dynamic Tooth Loads of Planetary Gear Sets having Tooth Profile Wear,” Mechanism and Machine Theory, 39, pp. 695-715.

Zhang, T., Kohler, H., and Lack, G., 1994, “Noise Optimization of a Double Helical Parallel shaft Gearbox,” International Gearing Conference, UK, pp. 93-98.

147

APPENDIX A

Beam Element Matrices

The formulations for the beam elements used to connect left and right sides of a gear sets are defined here. Considering an Euler beam element en (   srcp,,,1,, pN, n=1,2) having 6 degrees of freedom at each of its two nodes as shown in Figure A1, the displacement vector of the element is given as

y x  z  y  x  z left qen  . (A1) y x  z  y x  z right

Element mass matrix Men in Eq. (2.15b) is given as sum of three matrices such that

MMMMen e123 e e , (A2)

148

156  0156 00140Sym . 2 02204 2 220004 m 000000 Me1  , 420 54 0 0 0 13 0 156  054013000156 007000000140 22 01303 0 0 02204   22 1300030220004 000000000000

36  036 000Sym . 2 03 04 2 30004 Ime 000000 Me2  , 30 36 0 0 0 3 0 36   0360300036 000000000 22 030 000304  22 3000 030004 000000000000

0 00 000Sym .  0000 00000 J 000001 M  me . e3 0000000 3  00000000 000000000 0000000000  00000000000 000001/2000001

149

Here, m is the mass of the beam element per unit length, Ime and Jme are the diametral and polar mass moments of inertia of the element per unit length, and  is the length of the element. The stiffness matrix of the same element  en in Eq. (2.15a) is defined as

KKKKen e123 e e , (A3) where

12  012 000Sym . 2 06 04 2 60004 EIae 000000 Ke1  , 3 12 0 0 0 6 0 12   0120600012 000000000 22 0 6 02 0 0 0 6  04 22 60002060004 000000000000

 0   00   001Sym .   0000   00000  EA   K  000000 e2  0000000  ,   00000000   00 1000001   0000000000   00000000000     000000000000

150

0 00 000Sym . 0000 00000 GJ  K  ae 000001 e3 0000000 .  00000000 000000000 0000000000 00000000000  00000 1000001

Here, Iae and Jae are diametral and polar area moments of inertia, A is the cross- sectional area of the element, and E and G are the Young’s modulus and the shear modulus of elasticity, respectively. Similarly, gyroscopic matrix for the same element

 en used in Eq. (3.9a) is defined as

0  36 0 000Skew Sym .  3000 2 03040l  mIae 000000 . (A4) Gen  15A 03603000l   3600030360l 000000000 2 3000ll 03000 22 030ll 0003040 000000000000

151

y

 x y y x  y z z x x  z z

Area A

Figure A.1: An Euler beam element.

152

APPENDIX B

Overall System Matrices

Overall gear mesh matrix described in Eq. (2.17) is given as

KKKKmesh  m123 m m  (B1)

where

N  11 11 12  KKse11 () spi L K se 0 0 0  i1   21 22 11 12   KKKKse1122 se se se 0 0  N   0KKK021 22 () 11 0  se22 se spi R   i1  N  11 11 12   000KKKre11 () rpi L re   i1   21 22 11   000 KKKre112 re re  Km1   12   000 0Kre2   000 00    000 00  000 00    ()K00K021 () 21   sp11 L rp L   000 00   21  00K( sp1)R 00     

153

 0000    0000  0000    0000  12   K000re2   N  22 11 KKre2  () rpi R 0 0 0  i1    N  11 11 12   0KKKce11 () cpi L ce 0 Km2    i1   21 22 11 12   0KKKKce1122 ce ce ce   N  21 22 11  00KKKce22 ce () cpi R   i1    21  0K()cp1 L 00  0000  ()K00K21 () 21 rp11 R cp R   

 ()K0012   sp1 L   000  12   00K()sp1 R    12  ()K00rp1 L     000  12   00K()rp1 R   12  Km3   ()K00cp1 L     000  12   00K()cp1 R   11 22 22 22 12  KK+KKp11 e() sp 1 rp 1 cp 1 L K p 11 e 0   21 22 11 12   KKKKpe11 pe 11 pe 12 pe 12   21 22 22 22 22   0KKKKKp12 e p 12 e() sp 1 rp 1 cp 1 R     

154

Overall mass matrix described in Eq. (2.17) is given as

MMM  123 M (B2) where

(()MM11 M 12 0 0 se11 s L se 21 22 11 12 MMMMse1122 se se se 0  21 22 0MMM0se22 se () s R 11 000MMre1  () r L 21 000Mre1 000 0 M   1 000 0  000 0 000 0  000 0 000 0  000 0   

00 00  00 00 00 00  12 M000re1 22 11 12 MMre12 re M re 2 0 0 21 22 MMMre22 re () r R 0 0 M  11 12 2 00MMMce11 () c L ce 21 22 11 00MMMce112 ce ce 00 0M21 ce2 00 00  00 00 00 00    155

000 0  000 0 000 0  000 0 000 0  000 0  000 0 M   3 M00012 ce2  22 MMce2  () c R 0 0 0   11 12 0MMpe11 () p 1 L M pe 11 0   21 22 11 12 0MMMMpe11 pe 11 pe 12 pe 12   00MMM21 22  () pe12 pe 12 p 1 R   

Overall gyroscopic matrix as described in Eq. (3.10) is given as

GGG  123 G (B2) where

GG11 () G 12 0 0 se11 s L se 21 22 11 12 GGGGse1122 se se se 0  21 22 0GGG0se22 se () s R 11 000GGre1  () r L 21 000Gre1 000 0 G   1 000 0  000 0 000 0  000 0 000 0  000 0   

156

0000  0000 0000  12 G000re1 22 11 12 GGre12 re G re 2 00 21 22 GGG00re22 re () r R G2  0000  0000  0000 0000  0000 0000  

00 0 0  00 0 0 00 0 0  00 0 0 00 0 0  00 0 0  00 0 0 G3   00 0 0  00 0 0 0G11 () G G 12 0 pe11 p 1 L pe 11  21 22 11 12 0Gpe11 G pe 11 G pe 12 G pe 12  21 22 00 Gpe12 GG pe 12 () p 1 R     

In equations (B1), (B2) and (B3), sub-matrices for the first planet (i 1) are shown while they can be expanded to the sub-matrices for planets i (iN2, ) using the same structure provided in these equations.

157

APPENDIX C

Elements of the Forcing Vector

The forcing vector in Eq. (2.26) was given as

4N F=()tt Fk () (C1) k 1

It is formed by 4N components Fk ()t , each defined by a particular gear mesh based on its phasing relationship. The first four components that are associated with the meshes of the first planet (i 1) are defined as follows:

()L  0  fsp   0  0    f ()R  0  sp  0  0     0  0  0  0     F ()tket ()L () 0 F ()tket ()R () 0 1 sp sp1 , 2 sp sp1   , 0  0     0  0  ()L  0  fsp1     0  0   ()R  0 fsp1        

158

0  0     0  0  0  0     ()L 0 frp     0  0   ()R 0 frp     F ()tket ()L () 0 F ()tket ()R () 0 3 rp rp1 , 4 rp rp1   . 0  0     0  0  ()L  0  frp1     0  0   ()R  0 frp1       

Here, sub-vectors f are 6x1 vectors defined in Sections 2.2.1 and 2.2.2 and 0 is a zero vector of dimension 6. The remaining 4(N  1) terms corresponding to planets 2 to N can also be written in the same fashion.

159