Design Analysis and Control

download Design Analysis and Control

of 149

Transcript of Design Analysis and Control

  • 7/27/2019 Design Analysis and Control

    1/149

    Design, Analysis and Control of a

    Spherical Continuously Variable Transmission

    By

    Jungyun Kim

    Submitted in Partial Fulfillment of The

    Requirements for The Degree of

    Doctor of Philosophy

    in theSchool of Mechanical and Aerospace Engineering

    atSeoul National University

    February 2001

  • 7/27/2019 Design Analysis and Control

    2/149

    To My Parents

    i

  • 7/27/2019 Design Analysis and Control

    3/149

    Abstract

    This dissertation is concerned with the design, analysis, and control of a novel con-

    tinuously variable transmission, the spherical CVT (S-CVT). The S-CVT has a

    simple kinematic structure, infinitely continuously variable transmission character-

    istics, and transmits power via dry rolling friction on the contact points between

    a sphere and discs. The S-CVT is intended to overcome some of the limitations

    of existing CVT designs. Its compact and simple design and relatively simple con-

    trol make it particularly effective for mechanical systems in which excessively large

    torques are not required.

    We describe the operating principles behind the S-CVT, including a kinematic

    and dynamic analysis. A prototype is constructed based on a set of design spec-

    ifications and results of theoretical performance analysis. In order to provide a

    quantitative analysis of the spin loss of the S-CVT, which is one of the main sources

    of power loss, we develop an explicit formulation using a modified classical friction

    model, and an in-depth study of the velocity fields and normal pressure distribution

    on the contact regions. The proposed friction model includes the pre-sliding effect,

    i.e., Stribeck effects. Actual transmission ratios and power efficiency are obtained

    from experiments with a prototype testbench.

    The open-loop shifting system of the S-CVT reveals nonlinearity and unstable

    characteristics. In order to cancel the nonlinearity of the shifting system and to

    make it stable to shifting commands, we develop an input-state feedback controller

    based on exact feedback linearization. We also investigate the power efficiency of

    a generic dc motor, and present the results of a numerical investigation of the S-CVTs energy savings possibility benchmarked against a standard reduction gear.

    Furthermore, we develop a minimum energy control law for the S-CVT driven by a

    ii

  • 7/27/2019 Design Analysis and Control

    4/149

    dc motor, and present numerical simulation results that confirm the performance of

    the controller.

    Finally, we design and construct an S-CVT based mobile robot to realize the

    various advantages of the S-CVT into practical use. One of the key features of

    the mobile robot is the design of a novel pivot mechanism for planar accessibility.

    Results of both numerical simulations and experiments are presented to validate the

    robots performance advantages obtained as a result of using the S-CVT.

    Keywords: Continuously variable transmission; infinitely variable transmission;

    dry rolling friction; spin loss; feedback linearization; minimum energy control;

    mobile robot.

    iii

  • 7/27/2019 Design Analysis and Control

    5/149

    Contents

    Dedication i

    Abstract ii

    List of Tables viii

    List of Figures xi

    1 Introduction 1

    1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

    1.2 CVTs for Passenger Cars . . . . . . . . . . . . . . . . . . . . . . . . 6

    1.3 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

    1.4 Outline and Contributions . . . . . . . . . . . . . . . . . . . . . . . . 19

    2 Dynamic Analysis of the Spherical CVT 23

    2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

    2.2 Kinematics of S-CVT . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    2.2.1 Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    2.2.2 Operating Principles . . . . . . . . . . . . . . . . . . . . . . . 26

    2.3 Dynamics of S-CVT . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

    iv

  • 7/27/2019 Design Analysis and Control

    6/149

    2.3.1 Motion of Sphere . . . . . . . . . . . . . . . . . . . . . . . . . 31

    2.3.2 Shifting Dynamics . . . . . . . . . . . . . . . . . . . . . . . . 33

    2.4 Reaction Forces of the S-CVT . . . . . . . . . . . . . . . . . . . . . . 35

    2.4.1 Normal Reaction Force Exerted on the Variator: Fn . . . . . 36

    2.4.2 Shifting Reaction Force on the Sphere: D . . . . . . . . . . . 37

    2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

    3 Prototype Design and Experimental Results 39

    3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

    3.2 Issues in Mechanical Design . . . . . . . . . . . . . . . . . . . . . . . 41

    3.2.1 Normal Force Loading Device . . . . . . . . . . . . . . . . . . 41

    3.2.2 Capacity of Shifting Actuator . . . . . . . . . . . . . . . . . . 42

    3.3 Prototype Specifications . . . . . . . . . . . . . . . . . . . . . . . . . 43

    3.4 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

    3.4.1 Performance of S-CVT . . . . . . . . . . . . . . . . . . . . . . 47

    3.4.2 Strength and Life Prediction of S-CVT . . . . . . . . . . . . 49

    3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

    4 Slip Analysis of the Spherical CVT 53

    4.1 Friction Model Review . . . . . . . . . . . . . . . . . . . . . . . . . . 55

    4.2 Modified Friction Model for S-CVT . . . . . . . . . . . . . . . . . . . 59

    4.3 Spin Loss of the S-CVT . . . . . . . . . . . . . . . . . . . . . . . . . 60

    4.3.1 Velocity Fields on the Contact Surface . . . . . . . . . . . . . 60

    4.3.2 Normal Pressure Distribution . . . . . . . . . . . . . . . . . . 65

    4.3.3 Quantitative Analysis of Spin Loss . . . . . . . . . . . . . . . 66

    4.4 Slip Motion of the S-CVT . . . . . . . . . . . . . . . . . . . . . . . . 70

    4.4.1 Stick and Slip States . . . . . . . . . . . . . . . . . . . . . . . 70

    v

  • 7/27/2019 Design Analysis and Control

    7/149

    4.4.2 Slip Loss of the S-CVT . . . . . . . . . . . . . . . . . . . . . 70

    4.4.3 Slip Involved Contact Analysis . . . . . . . . . . . . . . . . . 70

    4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

    5 Shifting Controller Design via Exact Feedback Linearization 74

    5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

    5.2 Stability Analysis of S-CVT Shifting System . . . . . . . . . . . . . . 76

    5.3 Differential Geometric Preliminaries . . . . . . . . . . . . . . . . . . 78

    5.4 Shifting Controller Design via Input-State Linearization . . . . . . . 81

    5.4.1 Controllability and Linearizability . . . . . . . . . . . . . . . 82

    5.4.2 Input-State Linearization . . . . . . . . . . . . . . . . . . . . 83

    5.5 Shifting Controller Design . . . . . . . . . . . . . . . . . . . . . . . . 84

    5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

    6 Optimal Control of an S-CVT equipped Power Transmission 91

    6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

    6.2 Power Efficiency of a DC Motor . . . . . . . . . . . . . . . . . . . . . 936.2.1 DC Motor Dynamics . . . . . . . . . . . . . . . . . . . . . . . 93

    6.2.2 Power Efficiency of a DC Motor . . . . . . . . . . . . . . . . 95

    6.3 Investigation of S-CVT Energy Savings . . . . . . . . . . . . . . . . 96

    6.3.1 Control Design based on the Computed Torque Method . . . 98

    6.3.2 Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . 99

    6.4 Minimum Energy Control via a B-Spline Parameterization . . . . . . 101

    6.4.1 B-Spline Parameterization . . . . . . . . . . . . . . . . . . . . 102

    6.4.2 Gradients of the Objective Function and Constraint . . . . . 103

    6.4.3 Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . 105

    6.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

    vi

  • 7/27/2019 Design Analysis and Control

    8/149

    7 Case Study: An S-CVT based Mobile Robot 109

    7.1 Motivation for Mobile Robot Applications . . . . . . . . . . . . . . . 110

    7.2 MOSTS: An S-CVT Mobile Robot . . . . . . . . . . . . . . . . . . . 112

    7.2.1 Pivot Device for Planar Accessibility . . . . . . . . . . . . . . 112

    7.2.2 Prototype Design . . . . . . . . . . . . . . . . . . . . . . . . . 114

    7.3 Numerical and Experimental Results . . . . . . . . . . . . . . . . . . 115

    7.3.1 Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . 118

    7.3.2 Experimental Results . . . . . . . . . . . . . . . . . . . . . . 120

    7.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

    8 Conclusion 123

    References 125

    vii

  • 7/27/2019 Design Analysis and Control

    9/149

    List of Tables

    3.1 Specifications of prototype. . . . . . . . . . . . . . . . . . . . . . . . 46

    3.2 Endurance test condition. . . . . . . . . . . . . . . . . . . . . . . . . 51

    4.1 Maximal normal pressure comparison. . . . . . . . . . . . . . . . . . 66

    5.1 Candidates for k1, k2. . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

    6.1 Characteristic coefficients of dc motor. . . . . . . . . . . . . . . . . . 97

    6.2 Energy consumption; reduction gear vs. S-CVT. . . . . . . . . . . . 101

    6.3 Energy consumption with the minimum energy control. . . . . . . . 108

    7.1 Hardware specifications of general mobile robots. . . . . . . . . . . . 110

    7.2 DC motor charateristic coefficients of MOSTS. . . . . . . . . . . . . 115

    7.3 Energy consumption; MOSTS vs. differential drive. . . . . . . . . . . 121

    viii

  • 7/27/2019 Design Analysis and Control

    10/149

    List of Figures

    1.1 Classification of transmissions for vehicles. . . . . . . . . . . . . . . . 2

    1.2 Fuel consumption reduction for an engine. . . . . . . . . . . . . . . . 4

    1.3 Engine speed variation. . . . . . . . . . . . . . . . . . . . . . . . . . 5

    1.4 Classification of CVTs. . . . . . . . . . . . . . . . . . . . . . . . . . . 6

    1.5 Belts for belt drive CVT. . . . . . . . . . . . . . . . . . . . . . . . . 7

    1.6 Belt drive CVTs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    1.7 Variable stroke drives. . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    1.8 Full toroidal CVT, by courtesy of Torotrak. . . . . . . . . . . . . . . 10

    1.9 Structures for traction and friction drive CVT. . . . . . . . . . . . . 11

    1.10 Geometries of toroidal CVT, by courtesy of Torotrak and NSK. . . . 13

    1.11 Optimal operating line of an engine. . . . . . . . . . . . . . . . . . . 17

    1.12 Typical variogram. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

    2.1 Standard structure of S-CVT. . . . . . . . . . . . . . . . . . . . . . . 25

    2.2 Velocity constraint diagram. . . . . . . . . . . . . . . . . . . . . . . . 27

    2.3 Operating principles of S-CVT. . . . . . . . . . . . . . . . . . . . . . 28

    2.4 Ideal speed ratio of S-CVT. . . . . . . . . . . . . . . . . . . . . . . . 29

    2.5 Transmittable torque. . . . . . . . . . . . . . . . . . . . . . . . . . . 30

    2.6 Coordinate system and forces on S-CVT. . . . . . . . . . . . . . . . 31

    ix

  • 7/27/2019 Design Analysis and Control

    11/149

    2.7 Forces on variator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    3.1 3-dimensional concept view. . . . . . . . . . . . . . . . . . . . . . . . 40

    3.2 Normal force loading device using a spring. . . . . . . . . . . . . . . 42

    3.3 Schematic diagram of S-CVT. . . . . . . . . . . . . . . . . . . . . . . 44

    3.4 Assembly drawing of S-CVT. . . . . . . . . . . . . . . . . . . . . . . 45

    3.5 S-CVT prototype. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

    3.6 Testbench of S-CVT. . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

    3.7 Experimental results. . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

    3.8 Power efficiency of S-CVT. . . . . . . . . . . . . . . . . . . . . . . . 49

    3.9 Endurance test result of input disc. . . . . . . . . . . . . . . . . . . . 51

    4.1 Spin loss in traction drives. . . . . . . . . . . . . . . . . . . . . . . . 54

    4.2 Classical model of static, kinetic, and viscous friction. . . . . . . . . 55

    4.3 Pre-sliding displacement phenomenon. . . . . . . . . . . . . . . . . . 57

    4.4 Proposed friction model. . . . . . . . . . . . . . . . . . . . . . . . . . 60

    4.5 Contact of two bodies with different curvature. . . . . . . . . . . . . 61

    4.6 Velocity vector field on contact point. . . . . . . . . . . . . . . . . . 63

    4.7 Typical relative velocity vector diagram. . . . . . . . . . . . . . . . . 64

    4.8 Friction forces at the infinitesimal area of the contact surface. . . . . 67

    4.9 Spin losses on S-CVT at input speed of 3000 rpm. . . . . . . . . . . 69

    4.10 Dislocation of contact center. . . . . . . . . . . . . . . . . . . . . . . 71

    4.11 Change of normal pressure distribution in XZ plane. . . . . . . . . . 72

    5.1 Stability of the S-CVT shifting system. . . . . . . . . . . . . . . . . . 875.2 Tracking performance of the S-CVT shifting system. . . . . . . . . . 88

    5.3 Tracking error and corresponding control. . . . . . . . . . . . . . . . 88

    x

  • 7/27/2019 Design Analysis and Control

    12/149

    5.4 System behaviors of S-CVT during the gear ratio change. . . . . . . 89

    6.1 Diagram of an armature-controlled dc motor. . . . . . . . . . . . . . 93

    6.2 Efficiency of an armature-controlled dc motor. . . . . . . . . . . . . . 96

    6.3 Target profile of output speed. . . . . . . . . . . . . . . . . . . . . . 97

    6.4 Computed variator angle time profile. . . . . . . . . . . . . . . . . . 99

    6.5 Motor behaviors; reduction gear vs. S-CVT. . . . . . . . . . . . . . . 100

    6.6 Power consumption; reduction gear vs. S-CVT. . . . . . . . . . . . . 100

    6.7 Interpretation oftilde07Eg(p). . . . . . . . . . . . . . . . . . . . . . 104

    6.8 Optimal variator angle time profile. . . . . . . . . . . . . . . . . . . . 106

    6.9 Motor behaviors with the minimum energy control. . . . . . . . . . . 107

    6.10 Output behaviors with the minimum energy control. . . . . . . . . . 107

    7.1 Pivot device for planar accessibility of MOSTS. . . . . . . . . . . . . 112

    7.2 Electric circuit diagram of pivot switch and driving motor. . . . . . . 113

    7.3 Hardware prototype of MOSTS. . . . . . . . . . . . . . . . . . . . . 116

    7.4 The desired trajectory. . . . . . . . . . . . . . . . . . . . . . . . . . . 116

    7.5 Calculated wheel velocity profile. . . . . . . . . . . . . . . . . . . . . 117

    7.6 Trajectory of variator angle. . . . . . . . . . . . . . . . . . . . . . . . 118

    7.7 Motor behaviors of MOSTS. . . . . . . . . . . . . . . . . . . . . . . . 119

    7.8 Power consumption; MOSTS vs. differential drive. . . . . . . . . . . 120

    7.9 Experimental results. . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

    xi

  • 7/27/2019 Design Analysis and Control

    13/149

    Chapter 1

    Introduction

    Power transmissions are a universal element in nearly all mechanical systems, from

    a small-sized reduction gear in a compact disc drive, to a complex gear box (usually

    referred to as a transmission) in a vehicle. Although their components, sizes, and

    operating principles vary, their main objective is to effect changes in the sources

    power in the manner that corresponds to the load condition by manipulating the

    transmission ratio (or the gear ratio, i.e., the ratio of the input speed to output

    speed). Well-designed power transmissions eliminate the need for oversized power

    sources, and increase the power efficiency of overall the system. Even though power

    transmissions are required in various engineering fields, research activities are driven

    mainly by automobile manufacturers for their conventional transmissions. Thus, in

    this dissertation, an overview of power transmissions will be focused on automobile

    applications.

    1

  • 7/27/2019 Design Analysis and Control

    14/149

    1.1 Overview

    The generated power from ordinary power sources (internal combustion engines,

    electric motors, etc.) is much different from the necessary tractive force for driving

    vehicles. Hence, it is necessary to transform the power adequately from the source

    to the tractive force; the transmission of the vehicle takes this role. General trans-

    missions for vehicles can be primarily classified into manual transmissions (MTs)

    and automatic transmissions (ATs), according to its actuating mechanism for the

    shifting action (decision of shifting time, engaging/disengaging of the power flow el-

    ements, selecting the ratio, etc.). A detailed classification of transmissions is shown

    in Figure 1.1.

    A MT consists of dry clutch, which engages and/or disengages the power flow,

    a pair of synchronizing devices and constant meshing gear train for each gear ratio,

    and gear ratio selecting devices. Its structure and components are simple enough to

    allow for a considerable reduction in size and weight compared to a conventional AT.

    Transmission

    ContinuouslyVariable

    Transmission

    Figure 1.1: Classification of transmissions for vehicles.

    2

  • 7/27/2019 Design Analysis and Control

    15/149

    Moreover, a MT is built up with pure mechanical components and has no external

    power loss, such as a hydraulic system; thus the power efficiency is quite better than

    that of an AT.

    A conventional AT consists of wet clutches, planetary gear trains for each gear

    ratio, a hydraulic system for shifting action, an electro-hydraulic servo system for

    shifting control, and a torque converter. A torque converter is a unique device which

    has the multiple roles of a torque multiplication device, starting device, and torsional

    damper. Besides disadvantages in size and weight, a hydraulic system including a

    torque converter shows significant power loss, reducing the the overall efficiency of

    an AT (and ultimately the fuel economy of an AT equipped vehicle). However as

    the driving comfort of vehicle becomes the main concern, and greater effort is made

    toward improving the efficiency of ATs, the market share of AT equipped vehicles

    is growing rapidly.

    Power sources have complex efficiency characteristics according to driving con-

    ditions. For example, an internal combustion engine has different fuel consumption

    rates (or, brake specific fuel consumption: BSFC) according to its speed and torque

    while producing the same amount of power (see Figure 1.2). In this figure, there

    are two engine operating points which produce the same power for 120 km/hr with

    regard to different gear ratios. In the case of gear ratio A, which is greater than B,

    the BSFC value of this point is smaller than that of gear ratio B by 10%; hence one

    can conclude that a wide-spread of gear ratios is helpful for improving a fuel econ-

    omy. In addition, making more gear ratios can enhance the acceleration performance

    for the same reason. Many transmission engineers therefore endeavor to develop a

    transmission having more gear ratios. But making more gear ratios increases the

    size and weight of a transmission.

    The continuously variable transmission (CVT) has continued to be an object

    3

  • 7/27/2019 Design Analysis and Control

    16/149

    Engine

    torque

    Engine speed

    100 %

    110 %

    120 %

    130 %

    140 %

    150 %Constant power

    at 120 kph

    Driving resistancecurve B

    Driving resistance

    curve A

    BSFC curve

    Engine torquecurve at W.O.T.

    Vehicle speed 120 kph with gear ratio A

    Vehicle speed 120 kph with gear ratio B

    10% FC

    reduction

    Figure 1.2: Fuel consumption reduction for an engine.

    of considerable research interest within the mechanical design community, driven

    primarily by the automotive industrys demands for more energy efficient and en-

    vironmentally friendlier vehicles. Unlike conventional stepped transmissions (MTs

    and ATs), in which the gear ratio cannot be varied continuously, a CVT has a con-

    tinuous range of gear ratios that can, up to device-dependent physical limits, be

    selected independently of the transmitted torque. This feature of the CVT allows

    for engine operation at the optimum fuel consumption point consistent with the

    given output power requirements, thereby improving the engines power efficiency.

    Moreover, the CVT does not suffer from shifting shock (see Figure 1.3).

    In 1886, a CVT with rubber belt and pulleys made by Daimler Benz company

    was known as the first CVT to have been applied to a passenger car. About 1930,

    General Motors acquired the patent of the toroidal drive, which will be mentioned

    in a subsequent section in more detail, and tried to develop their own CVT. But

    they failed to commercialize it, and finished the related research in 1935. A different

    4

  • 7/27/2019 Design Analysis and Control

    17/149

    time

    speed

    time

    speed

    time

    speed

    Stepped

    Transmission

    CVT

    Vehicle

    Power source

    repeat accordingto the shifting

    stay aroundsome point

    regardless of shifting

    Figure 1.3: Engine speed variation.

    toroidal type CVT, known as a Heyes Self-Selector, was adopted in many Austin

    cars, although its production ceased after two years.

    The first commercially successful CVT for a passenger car was the rubber belt

    Variomatic of DAF Co., developed in 1958. The Variomatic was not popular, be-

    cause it failed to resolve the problems of rubber belt failure and the performance

    degradation due to deformation and wear. In the 1960s, a CVT using a metal belt

    and variable pulleys was developed by Hub Van Doorne, but did not make it to the

    market due to its insufficient torque capacity.

    In the 1970s, due to the worldwide oil-crisis and the raised environmental recogni-

    tion, many countries strengthened the regulations of the fuel economy and exhausted

    emissions of vehicles. Moreover by the advances of metallurgy and production tech-

    nology, inherent restraints of CVT could overcome; the research and development

    for CVT was much encouraged from the late of 1980s. Currently several automobile

    manufacturers have developed various prototype CVTs that are soon expected to

    5

  • 7/27/2019 Design Analysis and Control

    18/149

    appear in commercial vehicles (see [1]-[5] and references therein).

    1.2 CVTs for Passenger Cars

    According to the power transmission element and shifting mechanism, existing CVTs

    can be classified into belt drive, traction drive, variable stroke drive, and hydro-

    static/dynamic drive (see Figure 1.4).

    Belt Drive CVT

    In a belt drive CVT, a rubber or steel belt running on conically shaped variable

    diameter pulleys is used to transmit power at different drive ratios. According to

    the belt material, belt drive CVT can be divided into rubber, chain, and metal belt

    type. In Figure 1.5, the schematic diagrams of rubber, chain, and metal belt are

    shown. Because of their small power capacity, rubber belt CVTs are adopted in

    ContinuouslyVariable Transmission

    Friction Drive

    Figure 1.4: Classification of CVTs.

    6

  • 7/27/2019 Design Analysis and Control

    19/149

    (a) Rubber belt. (b) Chain belt. (c) Metal belt.

    Figure 1.5: Belts for belt drive CVT.

    compact cars and machine tools. Passenger cars equipped with a chain belt CVT

    had previously appeared on the market, but their production halted before long

    (a) ACVT of Aichi Co. (b) Multimatic of Honda Co.

    Figure 1.6: Belt drive CVTs.

    7

  • 7/27/2019 Design Analysis and Control

    20/149

    owing to chain noise and vibration problems.

    Currently almost conventional CVTs have push type metal belts of Van Doornes

    Transmissie b.v. or a revised form. Although the metal belt still suffers from a

    small torque capacity, the number of production units has been rising steadily in

    the worldwide market. The torque capacity has recently increased with the aid of

    advances in metallurgy and improvements in the hydraulic system. Figure 1.6 shows

    the rubber belt drive CVT made by Aichi Co. and metal belt drive CVT by Honda

    Co.

    Hydrostatic/Dynamic and Variable Stroke Drive CVT

    Hydrostatic/dynamic drives use an incompressible fluid as the transmission medium,

    by connecting a hydraulic pump directly to a variable displacement hydraulic actu-

    ator. It can realize neutral, forward, and reverse stages, but is not typically applied

    to passenger cars, owing to its own low power efficiency, weight, and noise. It has

    (a) Cylinder type. (b) Link type.

    Figure 1.7: Variable stroke drives.

    8

  • 7/27/2019 Design Analysis and Control

    21/149

    seen limited applications to heavy equipment.

    A variable stroke drive is made with one-way clutches and crank devices which

    can adjust the crank arm length. The rotational motion of the drive shaft transforms

    into translational motion, and the one-way clutch rectifies the motion into a uni-

    directional motion. This type of CVT cannot manage properly the pulsative output

    torques, and is therefore not adopted in vehicles (see Figure 1.7, by courtesy of DOE

    report [6]).

    Traction and Friction Drive CVT

    Friction wheels of unequal diameter were one of the earliest speed changing mech-

    anisms. It is speculated that their use even predates that of gearing toothed

    wheels, whose beginnings date back to the time of Archimedes, circa 250 B.C. [7].

    Even today, friction drives may be found in equipment where a simple and eco-

    nomical solution to speed regulation is required: phonograph drives, self-propelled

    lawnmowers, or even amusement park rides driven by a rubber tire are a few of the

    more common examples. In these examples, simple dry contact is involved, and thetransmitted power levels are low. However, this same principle can be harnessed

    in the construction of an oil-lubricated, all steel component transmission which can

    carry hundreds of horsepower using todays technology. In fact, oil-lubricated trac-

    tion drives have been in industrial service as speed regulators for more than 70 years

    [8].

    Great progress in tribology research since late 1960s, particularly research on

    elasto-hydrodynamic lubrication (EHL) traction, has made it easier to understand

    the traction drive mechanism. Traction drives transmit power through an increased

    shear force, which results from elasto-hydraulic shear stress of the traction oil be-

    tween two rotating solid bodies. A coefficient of traction is typically 0.1, and macro-

    9

  • 7/27/2019 Design Analysis and Control

    22/149

    scopic slip occurs at any time. Since there is no direct contact between the rotating

    bodies, wear phenomenon does not occur. The drive ratio is varied by controlling

    the effective onset radius of the contact point.

    According to the geometries of rotating elements, there are various type of trac-

    tion drive CVTs: nutating drive, half toroidal type, and full toroidal type. The half

    toroidal CVT has semi-circular discs, while full toroidal CVTs have full-circular discs

    as input/output rotating elements. According to the curvature radii of discs, they

    have different attainable gear ratios, torque capacity, and spin loss. Figure 1.8 shows

    the full toroidal drive and pertinent CVT made by Torotrak Ltd. in UK. Apart from

    this, many automobile manufacturers have developed half toroidal CVTs with dif-

    ferent torque capacities in Japan. Along with the traction oil developers (Santotrak,

    Shell companies), they have presented various prototypes of traction drive CVTs in

    the market [9].

    Generally, a traction drive shows rapid shifting response compared to belt drives,

    (a) Full toroidal traction drive. (b) Pertinent CVT.

    Figure 1.8: Full toroidal CVT, by courtesy of Torotrak.

    10

  • 7/27/2019 Design Analysis and Control

    23/149

    and can be adopted to medium-sized (even to large-sized) vehicles, because the

    highly-pressurized traction oil endures more shear stress than belt drives. The draw-

    backs of a traction drive are known to be as follows: the need for careful temperature

    control, sealing and supply of traction oil, and complicated shifting control due to

    the three dimensional contact curvature of the rolling elements.

    Finally, there exist friction drives where the power transmission mechanism is

    via rolling resistance and friction force in direct contact, though its structure and

    operating principle are much similar to traction drives (see Figure 1.9, by courtesy

    of DOE report [5]). Friction drives have been also found in several types of wood-

    working machinery dating back to before the 1870s. For example, [10] reports of a

    frictional gearing being used to regulate the feed rate of wood on machines in which

    one wheel was made of iron and the other, typically the driver, of wood (or iron

    covered with wood).

    Figure 1.9: Structures for traction and friction drive CVT.

    11

  • 7/27/2019 Design Analysis and Control

    24/149

    Friction drives have not been considered for passenger cars due to its low torque

    capacity, wear, heat dissipation problems, etc.However, friction drives have received

    significant attention from the perspective of tribology, because precise positioning

    can be accomplished while avoiding backlash [11]-[15]. Furthermore, traction and

    friction drives provide much design flexibility in terms of their structure and al-

    lowance for compact-sized designs.

    Although each type of CVT has its own particular set of advantages and disad-

    vantages, common difficulties shared by current CVTs are the complicated shifting

    controller design, and the need for a large-capacity, typically inefficient shifting ac-

    tuator [5]. Also, these CVT designs do not have infinitely variable transmission

    (IVT) capabilities, i.e., they do not include zero output speed among its available

    ratios, and therefore require a clutch or other type of starting and engaging device

    for initially driving the vehicle.

    1.3 Literature Review

    There is a vast amount of literature regarding the design, analysis, control, and

    application of CVTs in engineering fields. This section focuses on the areas of

    design and control of traction/friction drives, because the proposed spherical CVT

    in this thesis shows similar characteristics with respect to operating principles, power

    transmission and shifting mechanisms, and control laws.

    Traction Drive Designs

    It is well known that one of the earliest examples of the friction drive was the patent

    of C. W. Hunt in 1877 [16]. Basically the mechanism of that drive was a toroidal

    drive, which was developed for more than a decade thereafter. Applications of

    12

  • 7/27/2019 Design Analysis and Control

    25/149

    traction drives to automobiles have been studied since the beginning of this century.

    Prior to 1935, cars were called Friction Drive Cars, experimentally installed with

    such drives, had received some attention: it was widely believed that power was

    transmitted by friction between the rolling metallic elements.

    In the latter half of the 1960s, when elsto-hydraulic lubrication (EHL) became

    better understood [17], it was recognized that power was transmitted by traction.

    The performance of a traction drive depends to a large extent on the rheological

    properties of the fluid in the EHL contact [18]-[29]. In the 1970s, synthetic traction

    oil was developed which had a traction coefficient almost 50% higher than before,

    and practical use of the traction drive CVT was thought to be close at hand. It was,

    however, not realized, because the heat treatment of the rolling elements could not

    be achieved. A new type of traction oil being developed for automotive use shows

    some promise [30], [31].

    (a) Full toroidal CVT. (b) Half toroidal CVT.

    Figure 1.10: Geometries of toroidal CVT, by courtesy of Torotrak and NSK.

    13

  • 7/27/2019 Design Analysis and Control

    26/149

    As stated earlier, there are two main design streams in traction drives, for auto-

    motive use full and half toroidal type CVTs. A full toroidal CVT [32] has full-circular

    discs as input/output media and power rollers as shifting devices (see Figure 1.10

    (a)). In a full toroidal CVT, power rollers are located at the center of the toroidal

    shaped input and output discs. A hydraulic loading system has commonly been used

    to supply the normal force, which is necessary to transmit power via traction. Its

    shifting mechanism is based on the side-slip force generated by the velocity difference

    of the contact point.

    A half toroidal CVT uses semi-circular discs instead full-circular ones, though

    the shifting mechanism is not different from full toroidal CVTs (Figure 1.10 (b)).

    Many engineers including P. W. R. Stubbs (1980), Lubomyr O. Hewko (1986), M.

    Nakano (1991, 1999), H. Kumura (1999), and H. Machida (1999) have presented the

    trends and issues on half toroidal CVT designs for use in full-sized cars as a future

    driveline technology [33]-[38]. Nakano (1991) reported that the main reasons for the

    unsuccessful commercialization of toroidal CVTs were thought to be the inability to

    obtain sufficient performance with respect to the traction and viscosity performance

    of the traction fluid, the fatigue strength of the rolling elements, power transmission

    efficiency, transient ratio change controllability, and the issue of synchronization

    control in connection with the parallel arrangement of the traction elements [35].

    A traction drive CVT changes its speed ratio by controlling the side-slip force

    on the Hertzian contact area. Tanaka and Eguchi (1991) showed the principles of

    the speed ratio changing mechanism of half-toroidal CVTs and highlighted a digital

    compensation method for stabilization of the electro-hydraulically operated speed

    ratio control mechanism [39]. Fellows and Greenwood (1991) reported that it might

    not be possible to suppress hunting of the shifting control signal, depending on the

    control system used [40]. In addition to these results, there are many materials

    14

  • 7/27/2019 Design Analysis and Control

    27/149

    related to toroidal CVT controller designs (for example [41], [42], and references

    therein).

    When the ratio changes in a half toroidal traction CVT, the necessary contact

    force does not vary significantly compared to a full toroidal CVT [43]. Consequently,

    a loading cam system of a half toroidal CVT that produces contact force in pro-

    portion to the input torque can provide high efficiency over the entire speed ratio

    range, contrary to the hydraulic loading system of a full toroidal type. Moreover

    it has been reported that the full toroidal traction CVT suffers larger spin moment

    at the contact points than the half toroidal type, which tends to reduce its power

    capacity [44].

    The current design issues on toroidal type traction drives can be summarized as

    follows:

    the material of rolling elements is not sufficiently reliable because of high pres-sure and high temperature on the traction contact point;

    there is no affordable traction oil which satisfies all the conditions of automo-

    biles, although it has been reported that an adequate traction oil has been

    developed recently [9], [31];

    there are no bearings which can support high speeds and a large axial load;

    the normal force loading system (e.g., hydraulics or loading cam), which isnecessary to produce traction force, is inefficient and needs precise control for

    equalizing the normal forces on the contact points;

    there are difficulties on the control of transient ratio change and synchronizedprecise control in connection with the parallel arrangement of the traction

    elements.

    15

  • 7/27/2019 Design Analysis and Control

    28/149

    CVT Controls

    A CVT is originally intended to operate the power source in power efficient regimes

    by means of manipulating its gear ratio. Many previous efforts are focused on finding

    the power efficient regimes of sources and controlling the gear ratio of a CVT in order

    to run the source within those regimes. Here we review the previous analysis results,

    which describe ways of controlling a CVT for maximizing the fuel economy of a

    CVT-equipped vehicle as well as how to establish the shift schedule (the so called

    variogram, which describes the graphic relation between the engine and vehicle

    speeds) of a CVT for the vehicles performance objectives.

    Generally the power efficiency of a source is maximized at only one point over

    its operating region. In an internal combustion engine (see Figure 1.2), the fuel

    consumption is lowered for higher engine torque. On the other hand, it worsens for

    high engine speeds as the mechanical loss is large at those speed points. The pumping

    loss tends to be large for low engine speeds; hence, the fuel consumption also worsens

    for low engine speeds. These characteristics are consistent with theory. If we operate

    the engine only at the most efficient point, however, the driving performance may

    not be satisfied, because the driving torque to be generated for each vehicle speed is

    limited. Therefore, the control and optimization of automotive powertrain systems

    with a CVT is achieved by cooperative control of the engine and CVT (see [2]-[4],

    [45]). A drive-by-wire structure using an electric throttle control device is adopted

    for this engine consolidated CVT control [46].

    Figure 1.11 shows an optimal operating line (OOL) of a typical engine for max-

    imum fuel economy. Generally, an OOL is constructed simply by connecting staticBSFC contours through optimization. For improving the fuel economy of a vehicle,

    it is definitely helpful to control the CVT gear ratio so as to run the engine along

    16

  • 7/27/2019 Design Analysis and Control

    29/149

    100 %

    110 %

    120 %

    130 %

    140 %

    150 %

    Figure 1.11: Optimal operating line of an engine.

    this operating line. Most CVT-equipped vehicles use shift schedules (or variograms)

    in look-up table form, presetting the optimal gear ratios obtained from the static

    performance data of the engine and road tests of the prototype vehicle (see Figure

    1.12). However, this OOL does not involve the vehicle dynamics including accel-

    eration, performance objectives, because there is no consideration for the engine

    dynamics.

    The classical way to control CVT cars is to use some information on the gear

    ratio or on the transmitted torque which is then fed back by a PID controller [47]-[49].

    Only when using gain-scheduled controllers with typically 100 different gain points

    could the required performance be achieved. Kolmanovsky et al. (1999) explored the

    use of a CVT for torque management during mode transitions in lean burn gasoline

    engines [50]. They demonstrated that an intuitively sound CVT gear ratio control

    strategy which attempts to completely cancel the engine torque disturbance may

    result in unstable zero dynamics. They concluded the coordination of engine torque

    17

  • 7/27/2019 Design Analysis and Control

    30/149

    Figure 1.12: Typical variogram.

    production and CVT gear ratio control during mode transitions was mandatory.

    Takahashi (1998) proposed a scheme to minimize rate of fuel consumption by a

    direct fuel injection engine used by combination with CVT [45]. Target values for

    the engine and transmission which minimize fuel consumption ensuring driving per-

    formance were calculated based on the nonlinear optimization method. As a result

    of optimization, target values for air-fuel ratio and gear ratio were calculated and

    controlled by tracking. Because the calculation of partial differential was impossible

    at some operating points, he used a simplex method that did not require calculat-

    ing differential values. For minimization of fuel consumption function under various

    restrictions, penalty functions were also introduced.

    The non-minimum phase behavior of the CVT based powertrain system (without

    a torque converter) was mentioned in [51]. Considering this phase behavior of CVT,

    Guzzella and Schmid (1995) addressed an exact feedback linearization approach for

    a controller of CVT equipped vehicle [52]. In their works, the plant dynamics were

    18

  • 7/27/2019 Design Analysis and Control

    31/149

    exactly linearized over the complete operating range using feedback linearization.

    And as an application of the exact linearization approach, a kick-down controller

    was designed.

    1.4 Outline and Contributions

    This dissertation deals with the design, analysis, and control of a spherical CVT. A

    conceptual design of a particular spherical CVT (S-CVT) was proposed by Joukou

    Mitsusida in [53]. The S-CVT consists of a sphere, input and output discs, and

    variators. The rotating input and output discs are connected to the power source

    and output shafts, respectively, while the sphere is situated between the input and

    output discs. The transmission ratio is controlled by adjusting the location of the

    variator on the sphere, which in turn controls the axis of rotation of the sphere. It

    transmits power via dry rolling friction on the contact points of sphere and discs;

    therefore, there exists a torque limitation decided by the static friction force.

    The S-CVT, intended to overcome some of the aforementioned limitations of

    existing CVT designs, is marked by its simple kinematic design and IVT charac-

    teristics, i.e., the ability to transition smoothly between the forward, neutral, and

    reverse states without the need for any brakes or clutches. Moreover its relatively

    simple control makes it particularly effective for mechanical systems in which ex-

    cessively large torques are not required (e.g., mobile robots, household appliances,

    small-scale machining centers, etc.).

    In order to put the S-CVT to practical use, an analysis of its operating principles,

    power transmission and shifting mechanisms, and power capacity together with the

    consideration for issues of hardware design, needs to be performed. This dissertation

    is aimed at providing theoretical and practical solutions for these concerns, through

    19

  • 7/27/2019 Design Analysis and Control

    32/149

    an in-depth study of the design, dynamics, and control of the S-CVT. This work can

    be categorized into four parts:

    analysis of theoperating principles, kinematics, and dynamics in Chapter 2;

    hardware design issues including a slip loss analysis in Chapter 3, 4;

    shifting controller and minimum energy control law design in Chapter 5, 6;

    application for a wheeled mobile robot as a case study in Chapter 7.

    The subsequent achievements in this work can be exploited to the design and analysis

    of traction or friction drives having similar structure.

    Currently, we are carrying out the development of other S-CVT applications

    for small-capacity speed changers, e.g., bicycles, laundry machines, wind-propelled

    generating systems, potters spinning wheels, etc.There still remain several practical

    problems, such as realizing precise shaft alignments and increasing the torque capac-

    ity. Currently research efforts are being directed toward the application of traction

    fluid for the purpose of adopting the S-CVT for large torque capacity applications,e.g., hybrid vehicles, compact cars, etc.

    The detailed outlines and contributions of each chapter can be stated as follows.

    Operating Principles, Kinematics, and Dynamics

    Chapter 2 describes the conceptual design and operating principles of the S-CVT

    together with a detailed kinematic and dynamic analysis of its performance. In

    addition, there shows analytic interpretation for the reaction forces of S-CVT which

    are normally exerted on variator and discs, along with the definitions of their physical

    meanings.

    20

  • 7/27/2019 Design Analysis and Control

    33/149

    Hardware Design Issues and Slip Loss

    Chapter 3 presents the prototype specifications and a discussion of the main design

    issues, focusing on the normal force loading device and the shifting actuator capacity.

    Some experimental results are given on the actual transmission ratios and power

    efficiency obtained from a prototype testbench, to validate the operating principles

    and performance of S-CVT. We briefly address the strength and life estimation for

    the S-CVT, based on the well-known ball-bearing life theory.

    Spin loss of S-CVT, which is one of the main power losses of the S-CVT (and

    more generally for friction and traction drives) due to slippage, is formulated using

    a modified classical friction model in Chapter 4. The proposed friction model can

    involve pre-sliding effect i.e., Stribeck effects. For this, we perform an in-depth

    study of velocity fields and the normal pressure distribution generated on the contact

    regions. We also provide a quantitative analysis of the spin loss of the S-CVT. In

    addition, we discuss contact analysis involving slip, in which a shear force resulting

    from friction occurs on the contact surface.

    Shifting Controller Design and Minimum Energy Control

    The shifting system of the S-CVT has second-order nonlinear dynamics, for which

    typical open-loop control systems are likely to develop unstable characteristics. In

    order to cancel the nonlinearity of the S-CVT shifting system and to make it stable

    and responsive to shifting commands, we develop a feedback controller based on

    the exact feedback linearization method in Chapter 5. We first investigate the

    instability of the S-CVT shifting system using the Lyapunovs indirect method. Wethen present the input-state feedback controller design of the S-CVT shifting system,

    and investigate the stabilizing and tracking performance of the dedicated shifting

    21

  • 7/27/2019 Design Analysis and Control

    34/149

    controller by numerical simulation.

    Chapter 6 deals with a minimum energy control law for the S-CVT connected

    to a dc motor. We first investigate the general power efficiency of a dc motor. We

    then present the results of a numerical investigation of the S-CVT energy saving

    possibility benchmarked with a standard reduction gear. For this investigation, a

    computed torque control algorithm for the S-CVT is proposed. In addition, we

    describe a minimum energy control law of S-CVT connected to a dc motor. To do

    this, we describe the general power efficiency characteristics of a dc motor. Then

    the minimum energy control design is carried out via B-spline parameterization.

    Numerical results obtained from simulations illustrate the validity of our minimum

    energy control design.

    An S-CVT based Mobile Robot

    Finally, we propose an S-CVT based mobile robot (denoted as MOSTS for a Mobile

    rObot with a Spherical Transmission System) to realize the various advantages of

    the S-CVT, including the originally intended CVT characteristic of energy efficiency,into practical use in Chapter 7. In this chapter, we first address the motivation

    for applying the S-CVT to a wheeled mobile robot by first reviewing the current

    hardware designs of mobile robots and their power efficiency. We then present the

    hardware design of our S-CVT based mobile robot in accordance with the target

    performance. In addition, we propose a novel pivot mechanism which is necessary

    for planar accessibility using an internal gear and an uncontrolled dc motor. We

    perform both numerical simulations and experiments for various motion plans, in

    order to validate the realization of the robots operation, the CVT characteristics,

    and its energy saving possibility.

    22

  • 7/27/2019 Design Analysis and Control

    35/149

    Chapter 2

    Dynamic Analysis of the

    Spherical CVT

    2.1 Introduction

    In this chapter a new type of spherical continuously variable transmission (S-CVT)

    is described. The S-CVT, intended to overcome some of the aforementioned limi-

    tations of existing CVT designs, is marked by its simple kinematic design and IVT

    characteristics, i.e., the ability to transition smoothly between the forward, neutral,

    and reverse states without the need for any brakes or clutches.

    Because the S-CVT transmits power via rolling resistance between metal on

    metal, it has limitations on the overall transmitted torque, which is effectively de-

    termined by the static coefficient of friction and the magnitude of the normal forces

    applied to the sphere. Due to this torque limitation, the S-CVT is not intendedfor automobiles and other large capacity power transmission applications. Target

    applications for the S-CVT include mobile robots, household electric appliances,

    23

  • 7/27/2019 Design Analysis and Control

    36/149

    small-scale machine tools, and other applications with moderate power transmis-

    sion requirements. Although the current design of the S-CVT is based on friction

    drive designs, it is our expectation that the power capacity of the S-CVT can be

    increased by the use of traction oil, an issue which we do not pursue further in this

    dissertation.

    Other spherical CVT structures have been proposed for use in passive mobile

    robots and for use as nonholonomic joints in robot manipulators. Carl A. Moore et

    al. (1999) have reported a 3R passive robot, called the Cobot. The Cobot adopts a

    rotational CVT to provide smooth, hard virtual surfaces for passive haptic devices in

    place of conventional motors. Its rotational CVT consists of a sphere caged by four

    rollers, and adopts the joint speeds and task space speeds along with the steering

    angles as control inputs [54]. Another application can be found in underactuated

    manipulators, designed by Serdalen et al. (1994). This work proposes a new type

    of manipulator architecture using a CVT-type robot joint that takes advantage of

    the inherent nonholonomy of the CVT [55]. Although these systems are designed

    to manipulate the speed ratio using a CVT mechanism, their main purpose is not

    for power transmission to improve the energy efficiency. Furthermore, the shifting

    mechanism of the S-CVT is quite different from these previous designs, as will be

    described below.

    In this chapter, the conceptual design and operating principles of the S-CVT

    are described together with a detailed kinematic and dynamic analysis of its perfor-

    mance. Section 2 describes the basic kinematic structure and operating principles

    of the S-CVT. In Section 3, we examine the dynamics of the S-CVT by deriving

    the equations of motion and its shifting mechanism. Finally in Section 4, we exam-

    ine the reaction forces on the S-CVT, in particular those exerted normally on the

    variator and discs.

    24

  • 7/27/2019 Design Analysis and Control

    37/149

    2.2 Kinematics of S-CVT

    2.2.1 Structure

    The S-CVT is composed of three pairs of input and output discs, variators, and a

    sphere (see Figure 2.1). The input discs are connected to the power source, e.g., an

    engine or an electric motor, while the output discs are connected to the output

    shafts. The sphere, which is the main component of the S-CVT, transmits power

    from the input discs to the output discs via rolling resistance between the discs and

    the sphere. The variators, which are connected to the shifting controller, are incontact with the sphere like the discs, and constrain the direction of rotation of the

    sphere to be tangent to the rotational axis of the variator.

    Figure 2.1: Standard structure of S-CVT.

    25

  • 7/27/2019 Design Analysis and Control

    38/149

    The speed and torque transmission ratios of the S-CVT vary with the angular

    displacements of the variators; this will be described in further detail in the following

    subsection on the operating principles of the S-CVT. To transmit power from the

    discs to the sphere or from the sphere to the discs, a device that supplies a normal

    force to the sphere, such as a spring or hydraulic actuator, must be installed on each

    shaft. As can be seen in Figure 2.1, the structure and components of the S-CVT

    are simple enough to allow for a considerable reduction in size and weight compared

    to conventional transmissions. The orientations of the input and output shafts can

    also be located freely using rollers at arbitrary positions rather than discs.

    2.2.2 Operating Principles

    When the input device is actuated by a power source, the input disc rotates about the

    input shaft. This rotation in turn causes a rotation of the sphere, due to the condition

    of rolling contact without slip between the input discs and the sphere. Rotation of

    the sphere in turn causes a rotation of the output discs, and subsequently of the

    output shaft. In the absence of any contact between the sphere and the variator, theaxis of rotation of the sphere will largely be determined by an equilibrium condition

    among the various contact and load forces being applied to the sphere.

    The role of the variator is to control the axis of rotation of the sphere. Specifically,

    referring to Figure 2.2, the variator contacts the sphere at a point (marked by P)

    located directly above the sphere center. Since the variator rotates about an axis

    normal to the variator disc and passing through the variator center (marked by C1,

    C2), it follows that the contact point between the variator and the sphere undergoes

    a linear velocity in a direction tangential to the variator disc (marked by V1, V2).

    By adjusting the location of the variator (from C1 to C2) it is therefore possible to

    control the axis of rotation of the sphere (from 1 to 2); the axis will be parallel

    26

  • 7/27/2019 Design Analysis and Control

    39/149

    V1

    V2

    V

    1

    2

    P

    C2

    C1

    Figure 2.2: Velocity constraint diagram.

    to the line between the variator center and the sphere-variator contact point, and

    passing through the sphere center.

    By varying the axis of rotation of the sphere, it is in turn possible to vary the

    radius of rotation of the contact point between the input disc and the sphere, Ri,

    as well as the radius of rotation of the contact point between the output disc and

    the sphere, Ro (see Figure 2.3). In this way the speed-torque ratio of the S-CVT

    can be adjusted. Figure 2.3 shows the various alignments of the variator for the

    forward, neutral, and reverse states of the output shaft of the S-CVT. The neutral

    state, which corresponds to zero rotation of the output disc, is achieved when Ro

    becomes zero. As apparent from the figure, the forward, neutral, and reverse states

    can all be achieved by smoothly manipulating the variator alignment, without the

    need for any additional clutches or brakes.

    Assuming roll contact without slip, the speed and torque ratio between the input

    27

  • 7/27/2019 Design Analysis and Control

    40/149

    and output discs is related to the variator angle by the following relations:

    outin

    =riro

    tan (2.1)

    ToutTin

    =rori

    cot (2.2)

    where is the angular displacement of the variator, in and out are the respective

    angular velocities of the input and output shafts, Tin and Tout are the respective

    input and output torques, and ri and ro are the respective radii of the contact

    points of the input and output discs (see Figure 2.3). There are two design variables

    that prescribe the transmission ratio: the ratio of the input and output contact

    ri

    ro

    RRi

    Ro

    ri

    ro

    R

    Ri

    ri

    ro

    R

    Ri

    Ro

    variator angle

    Figure 2.3: Operating principles of S-CVT.

    28

  • 7/27/2019 Design Analysis and Control

    41/149

    -50-30

    -1010

    3050

    0.10.3

    0.50.7

    0.91.1

    -2.0

    -1.5

    -1.0

    -0.5

    0.0

    0.5

    1.0

    1.5

    2.0

    2.5

    Figure 2.4: Ideal speed ratio of S-CVT.

    radiiriro

    , and the variator angle . From Equation (2.1) it is apparent that a large

    range of available transmission ratios is possible even with a sphere of small radius.

    Assuming that there is no slip or other physical effects, the ideal speed ratio of the

    input speed to output speed is shown in Figure 2.4.

    Although ideally an infinite torque ratio is possible with the S-CVT as seen in

    Equation (2.2), in practice there is a limit to the torque that can be transmitted

    because power transmission occurs from rolling resistance of metal on metal. Figure

    2.5 shows a plot of the torque ratio as a function of the variator angle, for a given

    fixed input torque. The actual torque ratio of the S-CVT will lie somewhere in the

    operating region as indicated in the figure because of power loss due to friction, slip,

    heat generation. The limiting torque Tmax is determined by the static coefficient of

    friction s and the normal force N exerted by the output disc spring mechanism on

    29

  • 7/27/2019 Design Analysis and Control

    42/149

    the sphere according to the relation Tmax = rsN, where r is the contact radius of

    the disc. When either the input or output torque applied at the disc-sphere contact

    exceeds this limit, slippage can occur. Taking into account this limiting torque, the

    output torque for a given input torque Tin is given as follows:

    Tout = Tmax sat( TinTmax

    rori

    cot ) TLoss (2.3)

    where the saturation function sat() is defined by

    sat(x) =

    sgn(x) if|x|

    > 1

    x if |x| 1,

    and TLoss is the torque loss in S-CVT. Though assuming roll contact without slip

    (i.e., the speed ratio can be realized as the ideal case), the torque loss cannot be

    zero, because there exist some torque losses resulted from the spin moments and

    internal loads, etc., which will be discussed in Chapter 3 and 4.

    Figure 2.5: Transmittable torque.

    30

  • 7/27/2019 Design Analysis and Control

    43/149

    2.3 Dynamics of S-CVT

    2.3.1 Motion of Sphere

    To investigate the shifting mechanism of the S-CVT, we designate a reference frame

    XYZ situated at the center of the sphere, and a moving reference frame xyz, with

    z coinciding with the spin axis of the sphere (see Figure 2.6). The various external

    forces acting on the sphere are also shown in this figure, neglecting the normal

    forces exerted on the sphere-discs contact points to hold the sphere and the weights

    of sphere and discs.We define the driving forces which are delivered from the input discs as FZi1 and

    FZi2, and the reaction forces exerted by the load torque from the output discs as

    Figure 2.6: Coordinate system and forces on S-CVT.

    31

  • 7/27/2019 Design Analysis and Control

    44/149

    FZo1 and FZo2. Ftv1 and Ftv2 denote the forces generated by the shifting actuator

    acting at the sphere-variator contact points. The remaining reaction forces at the

    input and output discs and variators are respectively denoted by FXi1, FXi2, FY o1,

    FY o2, Fnv1, Fnv2.

    Assuming that the sphere center does not move and the rotational axis of the

    sphere lies on the xy plane, the force equilibrium conditions for each coordinate are

    as follows:

    FXo1 FXo2 + FXi1 FXi2 + (Fnv1 Fnv2)cos (Ftv1 Ftv2)sin

    FY i1

    FY i2 + FY o2

    FY o1 + (Fnv1

    Fnv2)sin + (Ftv1

    Ftv2)cos FZo1 FZo2 + FZi1 FZi2 + FZv1 FZv2

    = 0. (2.4)With respect to the specified coordinate frames, we can derive the dynamic equations

    relating the angular momentum change with the resultant moment acting on the

    sphere, i.e.,

    d

    dtHo =

    Mo

    where Ho is the angular momentum and Mo is the resultant moment. Theangular momentum of the sphere is given by:

    Ho = Is = 25

    msR21

    where is the angular velocity of the sphere, Is is its mass moment of inertia, ms

    the mass, and R the radius. Expressing the angular momentum of the sphere in

    terms of the moving coordinate frames, we obtain the derivatives of this momentum

    and the resultant moments, leading to the following set of second-order differential

    equations:

    Is

    =

    (Fnv1 + Fnv2)R (FZi1 + FZi2)R sin (FZo1 + FZo2)R cos

    (FY o1 + FY o2 + FXi1 + FXi2)R

    (Ftv1 + Ftv2)R (FZo1 + FZo2)R sin + (FZi1 + FZi2)R cos

    (2.5)

    32

  • 7/27/2019 Design Analysis and Control

    45/149

    where is the spinning rate of the sphere.

    Considering the attributes of the external forces, as stated earlier, we can restate

    those forces in Equation (2.5) as follows:

    FZi1 + FZi2 = Fi,

    FZo1 + FZo2 = Fo,

    Ftv1 + Ftv2 = Ft,

    Fnv1 + Fnv2 = Fn.

    In the above equations, Fn should not be regarded as an active force for shifting,

    but rather as a loss-like force acting to resist any variator displacements. Examining

    the reaction forces at the input and output discs caused by changes in the sphere

    axis of rotation, we can also conclude that the magnitudes of these forces must be

    equal, otherwise the sphere will be distorted:

    FXi1 = FXi2 = FY o1 = FY o2 = D. (2.6)

    The relevant forces can therefore be summarized as follows:

    Fi = Driving force delivered from the input discs;

    Fo = Reaction force caused by the output discs connected to the load torque;

    Ft = Shifting force on sphere delivered from the variator in the tangential direction;

    Fn = Loss-like reaction force exerted normally on variator;

    D = Reaction force on sphere generated by the shifting.

    2.3.2 Shifting Dynamics

    To establish the dynamic relations between the sphere and variator, we first define

    the forces on the upper sphere-variator contact point and the connected shifting

    33

  • 7/27/2019 Design Analysis and Control

    46/149

    Figure 2.7: Forces on variator.

    actuator (see Figure 2.7). To permit spinning motion of the variator, bearings arelocated on the connecting rod, which connects the shifting actuator and variator.

    In this figure, denotes the angular displacement of the shifting actuator, which

    consists of the same number of variators, m is the mass of the variator, and the

    eccentric distance between the centers of the shifting actuator shaft and variator. In

    addition, Fsv1 is the shifting force delivered by the shifting actuator, a is the linear

    acceleration of the variator center, and v1 the rotational speed of variator.

    Using the velocity constraint on the sphere-variator contact point, one can obtain

    the rotational speed of variator v1

    v1 = +R

    (2.7)

    34

  • 7/27/2019 Design Analysis and Control

    47/149

    where is the spinning rate of the sphere, and R the sphere radius as defined earlier.

    Let the moment of inertia of the shifting actuator shaft and connecting rod be Ia,

    and that of the variator be Iv. The shifting force delivered from the variator onto

    the sphere in the tangential direction (Ftv1) can be written as

    Ivv1 = Ftv1. (2.8)

    By the force relation Fsv1 = ma+Ftv1, and using the fact that the linear acceleration

    of the variator a = , as well as Equations (2.7), (2.8), we can express the shifting

    torque Fsv1

    ,

    Fsv1 = (Iv + m2 + Ia) + Iv

    R

    . (2.9)

    We assume that the lower variator always runs synchronously with the upper

    one; then the total shifting force Fs = 2Fsv1 and v1 = v2. Rearranging the

    equations of the sphere and variator (2.5) and (2.9), we obtain the following set of

    second-order differential equations for the S-CVT:

    2(Ia+Iv+m2)

    2RIv

    2

    2 Iv Is

    R + 2 RIv

    2

    = Fs

    Fi cos Fo sin . (2.10)The reaction forces are given by

    Fn =IsR

    + Fi sin + Fo cos , (2.11)

    D =1

    4

    IsR

    , (2.12)

    Ft = 2Iv

    ( +R

    ). (2.13)

    2.4 Reaction Forces of the S-CVT

    The two main reaction forces of the S-CVT are Fn and D, which are exerted respec-

    tively at the contact points between the sphere and discs, and sphere and variators.

    35

  • 7/27/2019 Design Analysis and Control

    48/149

    In order to prevent slippage, they must be smaller than the maximal friction force.

    In this section we derive analytic expressions for Fn and D, and examine their effect

    on the performance of the S-CVT.

    2.4.1 Normal Reaction Force Exerted on the Variator: Fn

    Fn, the reaction force which is exerted normally on the variators, can be considered

    as a loss force, and restricts the available gear ratios. Since Fn acts ultimately on

    the bearings located within the connecting rod, which connects the shifting actuator

    and variator (see Figure 2.7), it can therefore cause excessive bearing normal forces

    and bending moments on the variator and connected shafts.

    From Equation (2.11), Fn at steady state becomes

    Fn = Fi sin + Fo cos . (2.14)

    From the fact that the shifting effort Fs is zero at steady state, the relation between

    Fi and Fo of Equation (2.10) becomes

    Fi cos = Fo sin .

    Substituting Fi into Equation (2.14), Fn becomes

    Fn =Fo

    cos . (2.15)

    Beyond a certain variator angle, the magnitude of Fn becomes larger than the max-

    imal friction force which is determined by the static coefficient of friction s and

    the normal force N; slippage therefore occurs at the sphere-variator contact point

    (similar to the limiting torque Tmax).

    During transient states, the dynamics of sphere and variator influences the

    magnitude of Fn additionally. More than any other reaction forces on the S-CVT,

    36

  • 7/27/2019 Design Analysis and Control

    49/149

    Fn varies considerably together with input/output force and the shifting dynamics;

    thus it contributes the limit of available gear ratios of S-CVT. The allowable range

    of Fn during transient states is

    Fn =Fo

    cos +

    IsR

    sN .

    Rearranging this, we obtain a range for the gear ratio :

    | | cos1 FosN IsR

    . (2.16)

    In order to increase the range of available gear ratios, one can reduce the internal

    load and hence increase the output force Fo, or decrease the shifting response , as

    well as improve material properties with respect to s, N.

    2.4.2 Shifting Reaction Force on the Sphere: D

    There are four contact points between the sphere and input/output discs in the S-

    CVT (see Figure 2.1). When shifting (i.e., changes in gear ratio) occurs, the reaction

    force D, which resists the angular momentum change of the sphere, appears at each

    contact point. The reaction force D is normally exerted on the discs, and it acts

    directly on the bearings located within the input/output shafts; thus it can be

    considered as loss force like Fn.

    Moreover from Equation (2.12), D is related with the shifting response , and acts

    to restrict the available shifting response. As is the case for Fn, slippage resulting

    from a reaction force larger than the maximal friction force makes the S-CVT unable

    to transmit power; therefore the following inequalities must be hold:

    D sN , 4RIs

    sN . (2.17)

    From this relation, we can conclude that the shifting response is constrained by the

    material properties s, the sphere geometries R, Is, and the normal force N.

    37

  • 7/27/2019 Design Analysis and Control

    50/149

    2.5 Summary

    In this chapter, we have presented the design and analysis of a newly developed

    spherical continuously variable transmission (S-CVT) focusing on its basic structure

    and operating principles, shifting mechanism, and its reaction forces. The S-CVT is

    intended to overcome some of the limitations of existing CVTs, e.g., difficult shifting

    controller design, and the necessity of a large-capacity and typically inefficient shift-

    ing actuator. It is marked by its simple configuration, infinite variable transmission

    (IVT) characteristics and realization of forward, neutral, and reverse states without

    any brakes or clutches.

    Because the S-CVT transmits power through rolling resistance between metal on

    metal, torque limitations prevent current versions of the S-CVT from being applied

    to large capacity power transmission systems like passenger cars. However, our

    study suggests that it can be well-suited for applications involving small mechanical

    systems such as mobile robots, household electric appliances, small-scale machining

    centers, etc.

    Finally, we have investigated the reaction forces which are exerted normally on

    the variator and discs. Both Fn and D constitute sources of power loss for the

    S-CVT; in particular, the magnitude variation of Fn along the variator angle is

    steeper than any other forces on S-CVT. Moreover Fn can be a dominant factor in

    determining the available range of gear ratios of the S-CVT. The shifting reaction

    force D is related with the shifting response and acts to restrict the available

    shifting response.

    38

  • 7/27/2019 Design Analysis and Control

    51/149

    Chapter 3

    Prototype Design and

    Experimental Results

    3.1 Introduction

    In designing a transmission, one must consider both the power capacity of the trans-

    mission and power source as well as the load conditions. In this chapter, we first

    define the design objectives of the S-CVT, taking into account its inherent charac-

    teristics such as the power transmission mechanism based on friction force, shifting

    mechanism, and operating principles.

    The proposed S-CVT is intended for use in small capacity mechanical systems,

    e.g., mobile robots, household electric appliances, small-scale machine tools, and

    other applications with moderate power transmission requirements. In determining

    the hardware specifications of the S-CVT, practical issues such as the amount ofnormal force required to assure rolling resistance at the contact points of the S-

    CVT, and the capacity of the shifting actuator that can realize the desired shifting

    39

  • 7/27/2019 Design Analysis and Control

    52/149

    response, and the range of available gear ratios must all be considered. Based on the

    kinematic and dynamic analysis results of the previous chapter, we have designed

    and built the following S-CVT prototype.

    In this chapter, we present the prototype hardware specifications for the S-CVT,

    and an analysis of its performance. Using a prototype testbench, we obtain ex-

    perimental results that serve to validate the operating principles and performance

    of the S-CVT. In Section 2, we discuss various issues in the mechanical design of

    the S-CVT, focusing on the normal force loading device and the shifting actuator

    capacity. To assure rolling resistance force at the contact points of the S-CVT, we

    adopt compressible springs because of their simple structure and the ease in ad-

    justing the preset load. To determine the shifting actuator capacity, we derive the

    Figure 3.1: 3-dimensional concept view.

    40

  • 7/27/2019 Design Analysis and Control

    53/149

    numerical relationship between the necessary power and the shifting demand using

    the previous dynamic analysis results. Section 3 shows the hardware specifications

    and schematic drawings of the prototype S-CVT. In Section 4, we present experi-

    mental results on the actual transmission ratios and power efficiency obtained from

    the prototype testbench. Finally, we briefly address the strength and life estimation

    of the S-CVT, based on the well-known ball-bearing life theory.

    3.2 Issues in Mechanical Design

    Among the relevant issues in designing the S-CVT, we will focusing in particular

    on the normal force loading device and the capacity of the shifting actuator. The

    considered issues are mainly related to power capacity, namely the maximal trans-

    mittable force and the shifting actuator design.

    3.2.1 Normal Force Loading Device

    In order to assure rolling resistant force at the contact points of the S-CVT, an appro-

    priate normal force should be applied on the sphere and discs. Compressible springs

    are employed at each shaft, which are connected to the variators and input/output

    discs, to make the mechanical structure simple and to adjust the amount of normal

    force easily (see Figure 3.2). Since the spring force is closely related with the limit

    of transmittable force, we need to measure and adjust it. Using a set-screw, the

    amount of normal force can be adjusted by fixing the preset displacement of the

    spring. To set an accurate spring force, strain gauges are attached to each relevant

    shaft.

    Regarding the amount of normal force, the larger spring force increases the trans-

    mittable force. However, applying too large normal force causes yielding and plastic

    41

  • 7/27/2019 Design Analysis and Control

    54/149

    Figure 3.2: Normal force loading device using a spring.

    deformation of the sphere and discs; careful consideration for the normal stress on

    the contact regions must be carried out. In this study, we have designated the nor-

    mal force amount as 100 kgf using the corresponding finite element analysis results

    obtained by ANSYS.

    3.2.2 Capacity of Shifting Actuator

    In order to determine the capacity of the shifting actuator, it is necessary to inves-

    tigate the variation of shifting force Fs along with the desired performance. From

    Equation (2.10), in steady state Fs is zero and the input-output force relation be-

    comes Fi cos = Fo sin . To achieve shifting (i.e., gear ratio change), a non-zero Fs

    must be induced by the shifting actuator in an appropriate manner.

    For example, we consider the case when shifting occurs by the amount d at a

    42

  • 7/27/2019 Design Analysis and Control

    55/149

    certain steady state instant. At the beginning of shifting, we can assume that the

    input-output force relation still holds. Rearranging Equation (2.10), the shifting

    force Fs becomes

    Fs = 2 (Ia + Iv + m2)

    2 I

    2v R

    2

    (Is2 + 2 IvR2)

    d . (3.1)

    As seen in Equation (3.1), Fs necessary for shifting is determined by R, d, and

    the mass moments of inertia of the sphere Is, variator Iv, and connected elements

    Ia + m2. Considering that shifting forces of other traction or belt drives must be

    large enough to resist the traction or friction force, which is generated directly by

    the transmitted torque, the overall magnitude of shifting force of the S-CVT will

    likely be much smaller than that of other existing CVTs.

    The necessary power Ps of the shifting actuator is calculated using Equation

    (3.1):

    Ps = Fsd = 2

    (Ia + Iv + m2) 2I

    2v R

    2

    (Is2 + 2 IvR2)

    d d (3.2)

    where d is the corresponding angular velocity to the required shifting demand d.

    3.3 Prototype Specifications

    Based on the numerical investigation results from the previous studies, we have des-

    ignated the hardware specifications of the S-CVT prototype. The overall layout of

    the power transmission is shown in Figure 3.3. Because of the maximum limiting

    torque, a reduction gear with a ratio of three is added to the prototype; this ratio

    also includes a safety factor. This additional reduction gear can be eliminated by

    improving the material properties such as the static coefficient of friction and in-

    creasing the normal force at the contact point. The final assembly drawing is shown

    43

  • 7/27/2019 Design Analysis and Control

    56/149

    Driving Motor

    PivotMotor

    z=36z=18

    z=18z=18

    z=36

    z=

    72

    z=

    36

    z=

    26

    z=

    52 z

    =26

    z=

    52

    z=

    26

    z=

    39z

    =26

    z=

    39

    Variator

    Sphere

    Wheel

    Wheel

    10

    VariatorMotor

    z=18 z=66

    z=66z=18

    13

    13

    Input disc

    Figure 3.3: Schematic diagram of S-CVT.

    in Figure 3.4. In Figure 3.3 and 3.4, a dc motor referred to as the pivot motor, and

    internal gears are included in the power-flow line of the S-CVT. These elements are

    added in order to make each output shaft rotate in opposite directions. This novel

    pivot mechanism is proposed for the application to the CVT-based mobile robot,

    which is the main subject of Chapter 7.

    44

  • 7/27/2019 Design Analysis and Control

    57/149

    load spring

    housing

    Figure 3.4: Assembly drawing of S-CVT.

    45

  • 7/27/2019 Design Analysis and Control

    58/149

    Element Nomenclature and Specification Material

    Spheremass (ms) = 0.882 kg

    radius (R) = 30 mm

    Steel ball

    of ball bearing

    Input/output disc

    Variator

    mass (m) = 0.095 kg

    radius (r) = 16 mm

    contact radius (ri/o, ) = 10 mm

    SNCM 8 class

    Input/output shafts SCM 4 class

    Gears Refer to Figure 3.3 SCM 21 class

    Mass moments

    of inertia

    Sphere (Is) = 3.1758 104 kg m2

    Input parts (Iin) = 2.3581 105 kg m2

    Output parts (Iout) = 3.8609 104 kg m2

    Variator (Iv) = 1.0514 105

    kg m2

    Variator connected parts (Ia) = 1.0585 104 kg m2

    Table 3.1: Specifications of prototype.

    The detailed specifications for numerical studies and experiments are shown in

    Table 3.1. The prototype S-CVT has been built and is shown in Figure 3.5.

    Figure 3.5: S-CVT prototype.

    46

  • 7/27/2019 Design Analysis and Control

    59/149

    3.4 Experimental Results

    In order to validate the operating principles and performance of the S-CVT, we have

    built a testbench for it. Two eddy-current type AC servo motors (input: 3-phase

    AC, 122 V, 9 A; output: 1500 Watts; rated speed: 2000 rpm) are used for a driving

    power source and a driven load generator. In the testbench (see Figure 3.6), the

    variator angle is controlled by a dc stepped motor with an angular resolution of

    0.024/pulse. The rotational speeds of the input and output shafts are measured

    through incremental optical encoders attached to the shafts.

    3.4.1 Performance of S-CVT

    Setting the external load torque to zero, we observe the output speed together with

    the variator angle displacement while the input speed is set respectively to 748,

    Figure 3.6: Testbench of S-CVT.

    47

  • 7/27/2019 Design Analysis and Control

    60/149

    1502, and 2001 rpm. A steady-state speed ratio curve of the S-CVT is extracted

    for the no-load condition (see Figure 3.7 (a)). Note that the overdrive of the output

    speed, which implies that the output speed is faster than the input speed, occurs

    when the variator angle exceeds 50. In addition, there is a large deviation between

    the ideal value and the test result beyond a variator angle of 65, which indicates

    the onset of slippage. These less than ideal output speeds arise from the increase of

    reaction force normally exerted on variator Fn, which is described in Section 2.4. In

    the experimental result, moreover, there must be a certain amount of internal load

    induced by manufacturing and other errors, which makes Fo large (see more details

    in Section 2.4).

    Using slip-ring type torque sensors, we have also observed the output torque

    together with the variator angle displacement by adjusting input/output torque to

    realize the pre-obtained steady state speed ratio (see Figure 3.7 (b)). The actual

    torque ratio is limited to under 20, which is determined mainly by the static coeffi-

    0 10 20 30 40 50 60 70

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    1.4

    1.6

    1.8

    (a) Speed ratio of S-CVT.

    0 10 20 30 40 50 60 70

    -5

    0

    5

    10

    15

    20

    25

    30

    35

    40

    45

    (b) Torque ratio of S-CVT.

    Figure 3.7: Experimental results.

    48

  • 7/27/2019 Design Analysis and Control

    61/149

    0 10 20 30 40 50 60

    0

    10

    20

    30

    40

    50

    60

    70

    80

    90

    100

    Figure 3.8: Power efficiency of S-CVT.

    cient of friction and the exerted normal force.

    Finally, we calculate the power efficiency of the S-CVT using the obtained speed

    and torque ratios (see Figure 3.8). The efficiency is almost 85% for variator an-

    gles under 15, while the average efficiency beyond this angle is about 65%. The

    power efficiency of the prototype S-CVT is somewhat low; this is mainly due to the

    manufacturing errors including bearing friction loss, gear backlash, etc. From exper-

    iments with the prototype S-CVT, we have also found that slight misalignments of

    the shafts may cause bending moments in the shafts and discs, resulting in increased

    bearing friction loss and slippage, although for applications accurate shaft alignment

    will have to be separately addressed.

    3.4.2 Strength and Life Prediction of S-CVT

    Perhaps the most common form of mechanical failure in friction and traction drives

    is by wear. The laws governing the overall friction and wear between two surfaces

    49

  • 7/27/2019 Design Analysis and Control

    62/149

    seem to depend primarily on the total force transmitted across the two surfaces

    rather than on the local distributions of stress and strain, and one may assume that

    the two bodies in contact are perfectly rigid. In studying the details of the actual

    mechanism of wear and friction, however, one must take into account the extremely

    small areas of actual load contact between two bodies and the elastic and plastic

    deformations in these regions. Another factor which must be considered in studying

    the detailed mechanism is the surface condition of the metal, since this condition

    may be such that these local points of contact behave in a manner quite different

    from that of the same material in bulk form. It has been generally accepted that

    the addition of a reasonable tangential force to a rolling contact has no appreciable

    effect on drive life. This is so only when spin is almost entirely absent.

    Dawe and Lohr (1993) reported that application of a realistic tangential trac-

    tion force at the contacts does not seem to cause dramatic reduction in life, and

    circular contacts appear to offer the best