Sei sulla pagina 1di 224

FINITE ELEMENT ANALYSIS OF A

SHAFT-ROTOR SYSTEM

Donald Andrew Phillips

Thesis submitted to the faculty of the Virginia Polytechnic and State University in Partial
fulfillment of the requirements for the degree of

Master of Science
In
Engineering Science and Mechanics

Romesh Batra, Co-chair


Norman Dowling, Co-chair
Steve Kampe

January 23, 2001


Blacksburg, Virginia

Keywords: Steady State Creep, Finite Element Method, Rotor, High Stress, High Temperature,
More Electric Aircraft
FINITE ELEMENT ANALYSIS OF A SHAFT-ROTOR SYSTEM

by
Donald Andrew Phillips

(ABSTRACT)
The United States Air Force is in the process of developing a more electric aircraft. The
development of an aircraft Integrated Power Unit and an Internal Starter/Generator will be
instrumental in producing sufficient electrical power to run all non propulsive systems. Iron-
cobalt alloys, such as Hiperco alloy 50HS, are high temperature, high strength magnetic
materials ideal for these power applications. Design requirements and previous studies indicate
that these materials need to survive in temperatures up to 1000°F (810K), rotation speeds of
about 55,000 rpm, and have strengths in excess of 80 ksi. Research conducted by Fingers
provided the material and creep properties used in the analysis presented in this report. The
finite element method was used to analyze a spinning rotor mounted to a circular shaft via an
interference fit subjected to various operating environments. The power law creep model
defined by Fingers was used to analyze three distinct rotor configurations. The first
configuration was a constant temperature single lamina, mounted to a shaft of equal thickness,
subject to temperatures between 727K and 780K, rotation speeds between 35,000 rpm and
60,000 rpm, and two different interference fits: 0.0015 inches and 0.003 inches. The results
yield conservative predictions that indicate that these models could not survive the required
operating conditions. The second configuration was a linear radial variation in temperature
single lamina, mounted to a shaft of equal thickness, subjected to three temperature ranges,
rotation speeds between 30,000 rpm and 55,000 rpm, and two different interference fits; 0.0015
inches and 0.003 inches. These results represent a more realistic model, which indicate that the
“cooler” inner portions of the rotor restrict the creep deformations of the “hotter” outer portions
resulting in higher possible operating temperatures and rotation speeds very near the required
operating conditions. The third configuration was a lamina stack comprised of two rotor lamina,
with a Coulomb friction surface interaction, and held together by a compressive axial force.
These models represent a first step towards understanding the behavior of the entire rotor stack.
ACKNOWLEDGMENTS

Drs. Romesh Batra, Norman Dowling, and Steve Kampe for all their help, knowledge, and
insights.

Dr. Richard Fingers for his support, encouragement, and his continued confidence that there was
a light at the end of the tunnel.

Loretta Tickle for keeping an eye out for me and without whom I could never have navigated my
way through all of the paperwork and regulations involved in pursuing my education. Whatever
they are paying her it is not nearly enough.

The “Aerospace Crew” for providing the comic relief necessary to maintain sanity during this
stressful time in my life.

Ann Marie and Carolyn Phillips, my sisters. I am truly blessed to have two sisters who are the
two most amazing people I have the privilege to know, I only hope that I can be as good of a
brother for them as they are sisters for me.

MacDonald and Inés Phillips, my parents. There are no words to describe my thanks and
appreciation of them, without whom I would not be the person I am today.

This work was managed by the University of Dayton Research Institute, with financial support
by the Air Force Research Laboratory. Dr. Richard Fingers was the technical monitor of the
project.

iii
TABLE OF CONTENTS

1 INTRODUCTION AND LITERATURE REVIEW........................................................................1

1.1 MOTIVATION ....................................................................................................................................1


1.2 CREEP PHENOMENA .........................................................................................................................3
1.3 FINITE ELEMENT METHOD ...............................................................................................................9
1.4 PREVIOUS WORK ............................................................................................................................14
1.5 MATERIAL ......................................................................................................................................18
1.6 PROBLEM STATEMENT ...................................................................................................................20
1.7 APPROACH ......................................................................................................................................20

2 ANALYSIS PROCEDURE..............................................................................................................23

2.1 MESH GENERATION AND LOAD CALCULATIONS ...........................................................................23


2.2 FINITE ELEMENT ANALYSIS ...........................................................................................................33
2.3 POST PROCESSING OF RESULTS ......................................................................................................35

3 RESULTS AND DISCUSSION.......................................................................................................36

3.1 RESULTS FROM THE ELASTIC ANALYSIS ........................................................................................38


3.1.1 Single Lamina Results ...............................................................................................................40
3.2 RESULTS FROM THE CREEP ANALYSIS ...........................................................................................54
3.2.1 Single Lamina Results – Constant Temperature Models ..........................................................55
3.2.1.1 727K.................................................................................................................................................................55
3.2.1.2 760K.................................................................................................................................................................65
3.2.1.3 780K.................................................................................................................................................................74
3.2.2 Single Lamina Results – Radial Variation in Temperature Models..........................................81
3.2.2.1 588K – 810K ....................................................................................................................................................84
3.2.2.2 727K – 810K ....................................................................................................................................................88
3.2.2.3 810K – 588K ....................................................................................................................................................92
3.2.3 Multiple Lamina Stack Results - Axial Variation in Temperature Models ..............................95
3.2.3.1 The 45,000 rpm Models ...................................................................................................................................98
3.2.3.2 The 55,000 rpm Models .................................................................................................................................116

4 CONCLUSIONS/SUMMARY ......................................................................................................135

4.1 ACCOMPLISHMENTS .....................................................................................................................135


4.1.1 The Finite Element Model.......................................................................................................135
4.1.2 Creep Results ..........................................................................................................................136

iv
4.2 APPLICATIONS ..............................................................................................................................137
4.3 FUTURE RESEARCH ......................................................................................................................138

5 REFERENCES ...............................................................................................................................139

6 APPENDIX A..................................................................................................................................143

6.1 INTRODUCTION........................................................................................................................143
6.2 INPUT FILES ..............................................................................................................................144
6.2.1 Type=1 or 2 Input Files ..........................................................................................................145
6.2.1.1 Heading and Time Step Variables ..................................................................................................................145
6.2.1.2 Lamina Variables............................................................................................................................................146
6.2.1.3 Pole Variables.................................................................................................................................................151
6.2.1.4 Shaft Variables ...............................................................................................................................................152
6.2.1.5 Load Variables ...............................................................................................................................................153
6.2.2 Type = 3 Input File .................................................................................................................154
6.2.2.1 Heading and Interference Fit Variables ..........................................................................................................154
6.2.2.2 Lamina Variables............................................................................................................................................155
6.2.2.3 Pole Variables.................................................................................................................................................156
6.2.2.4 Shaft Variables ...............................................................................................................................................157
6.3 RUNNING GENMESH.F ..................................................................................................................158
6.4 RUNNING ABAQUS .....................................................................................................................159
6.5 RUNNING POSTP.F AND CREATING RESULTS FILES.......................................................................160
6.6 APPENDIX .....................................................................................................................................162
6.6.1 Appendix I – Example Type = 1 and 2 input file (input).........................................................162
6.6.2 Appendix II – Example ABAQUS input file (model.inp) .........................................................163
6.6.3 Appendix III – Example Heading Cards .................................................................................166
6.6.4 Appendix IV – Example Lamina Cards ...................................................................................167
6.6.5 Appendix V – Example Pole cards ..........................................................................................168
6.6.6 Appendix VI – Example Shaft Cards .......................................................................................168
6.6.7 Appendix VII – Example Load card ........................................................................................168
6.6.8 Appendix VIII – Example Type = 3 Input File (input) ............................................................169

7 APPENDIX B – CREEP TYPE A RESULTS..............................................................................170

8 APPENDIX C – CREEP TYPE AB1 RESULTS.........................................................................194

9 VITA ................................................................................................................................................214

v
TABLE OF FIGURES

Figure 1-1 – Integrated Power Unit Schematic ..........................................................................................................................................................2


Figure 1-2 – Typical Rotor/Stator Configuration [5] .................................................................................................................................................4
Figure 1-3 – Typical Creep Strain vs. Time Curve ....................................................................................................................................................4
Figure 1-4 – Estimation of the Circumference of a Circle........................................................................................................................................12
Figure 2-1 – Typical Eight Noded Finite Element ...................................................................................................................................................24
Figure 2-2 – Shaft-Rotor Geometry .........................................................................................................................................................................24
Figure 2-3 – Shaft-Rotor Mesh Generation Variables..............................................................................................................................................25
Figure 2-4 – Forces and Boundary Conditions Acting on the Shaft-Rotor System ..................................................................................................27
Figure 2-5 - Centrifugal Force acting on a Single Element ......................................................................................................................................27
Figure 2-6 - Node Numbers and interpolation functions for a Lagrange rectangular element..................................................................................28
Figure 2-7 - Integration (Gauss) Points for a Quadrilateral Element........................................................................................................................30
Figure 2-8 Radial and Tangential Stress Profiles for Internally and Externally Pressurized Cylinders ...................................................................31
Figure 2-9 – Calculating r “new” .............................................................................................................................................................................34
Figure 2-10 – Interference Fit Stress Profile ............................................................................................................................................................34
Figure 3-1 – Elastic Modulus vs. Temperature for HA50HS ...................................................................................................................................37
Figure 3-2 – 2400 Element Mesh.............................................................................................................................................................................39
Figure 3-3 – Stress Coefficient vs. Temperature for HA50HS.................................................................................................................................39
Figure 3-4 – Elastic Stress Profiles for 727K and 810K for d=0.0015’’, ω=45,000rpm ..........................................................................................41
Figure 3-5 – Elastic Radial displacement Profiles for 727K and 810K for d=0.0015’’, ω=45,000rpm....................................................................41
Figure 3-6 – Elastic Von Mises Strain Fringe Plot for d=0.0015’’, T=727K, ω=45,000rpm ...................................................................................42
Figure 3-7 – Elastic Von Mises Strain Fringe Plot for d=0.0015’’, T=810K, ω=45,000rpm ...................................................................................42
Figure 3-8 – Elastic Stress Profiles for 45,000 rpm and 55,000rpm for d=0.0015’’, T=727K .................................................................................44
Figure 3-9 – Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=45,000rpm, T=760K ..............................................................................44
Figure 3-10 – Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=45,000rpm, T=780K.............................................................................45
Figure 3-11 – Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=45,000rpm, T=795K.............................................................................45
Figure 3-12 – Elastic Pole-tip Radial displacement vs. Temperature for ω=45,000rpm ..........................................................................................47
Figure 3-13 - Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=55,000rpm, T=760K .............................................................................47
Figure 3-14 - Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=55,000rpm, T=780K .............................................................................48
Figure 3-15 - Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=55,000rpm, T=795K .............................................................................48
Figure 3-16 - Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=55,000rpm, T=810K .............................................................................49
Figure 3-17 - Elastic Stress Profile for Shaft-Hub System d=0.003’’, ω=55,000rpm, T=727K ...............................................................................49
Figure 3-18 - Elastic Stress Profile for Shaft-Hub System d=0.003’’, ω=55,000rpm, T=760K ...............................................................................50
Figure 3-19 - Elastic Stress Profile for Shaft-Hub System d=0.003’’, ω=55,000rpm, T=780K ...............................................................................50
Figure 3-20 - Elastic Stress Profile for Shaft-Hub System d=0.003’’, ω=55,000rpm, T=810K ...............................................................................51
Figure 3-21 – Elastic Pole-tip Radial displacement vs. Temperature for d=0.003’’, ω=55,000rpm.........................................................................53
Figure 3-22 – Elastic Pole-tip Radial displacement vs. RPM for d=0.003’’, T=727K .............................................................................................53
Figure 3-23 – Shaft-Hub Radial Stress profile for d=0.0015’’, ω=50,000rpm, T=727K, Type A, ATi=AFeCo ......................................................57
Figure 3-24 – Shaft-Hub Radial displacement Profile for d=0.0015’’, ω=50,000rpm, T=727K, t=75hrs, Type A, ATi=AFeCo.............................57
Figure 3-25 – Von Mises Strain Fringe Plot for d=0.0015’’,ω=50,000rpm, T=727K, t=75hrs, Type A, ATi=AFeCo ............................................58
Figure 3-26 – Radial displacement vs. Time for d=0.0015’’, ω=50,000rpm, T=727K, Type A, ATi=AFeCo.........................................................58
Figure 3-27 – Pole-tip Radial displacement vs. RPM for d=0.0015’’, T=727K, Type A, ATi=AFeCo ...................................................................59
Figure 3-28 – Pole-tip Radial displacement vs. RPM for T=727K, Type A, d=0.0015’’ and d=0.003, ATi=AFeCo’’ ............................................61

vi
Figure 3-29 – Shaft-Hub Radial Stress Profile for d=0.003’’, T=727K, ω=60,000rpm, Type A, ATi=AFeCo........................................................61
Figure 3-30 – Shaft-Hub Radial displacement Profile for d=0.003’’, ω=60,000rpm, T=727K, t=2292hrs, Type A, ATi=AFeCo...........................62
Figure 3-31 – Von Mises Strain Fringe Plot for d=0.003’’,ω=60,000rpm, T=727K, t=2292hrs, Type A, ATi=AfeCo ...........................................62
Figure 3-32 – Radial displacement vs. Time for d=0.003’’, T=727K, ω=60,000rpms, Type A, ATi=AFeCo .........................................................64
Figure 3-33 – Pole-tip Radial displacement vs. RPM for T=727K, Type AB1, d=0.0015’’ and d=0.003’’, ATi=AfeCo ........................................64
Figure 3-34- Shaft-Hub Stress Profile for d=0.0015’’, T=760K, ω=40,000rpm, Type A, ATi=AFeCo...................................................................67
Figure 3-35 – Shaft-Hub Radial displacement Profile for d=0.0015’’, ω=40,000rpm, T=760K, t=2052hrs, Type A, ATi=AFeCo.........................68
Figure 3-36 – Von Mises Strain Fringe Plot for d=0.0015’’,ω=40,000rpm, T=760K, t=2052hrs, Type A, ATi=AFeCo.........................................68
Figure 3-37 – Radial displacement vs. Time for d=0.0015’’, T=760K, ω=40,000rpms, Type A, ATi=AfeCo .......................................................69
Figure 3-38 – Pole-tip Radial displacement vs. RPM for T=727K and T=760K, Type A, d=0.0015’’, ATi=AfeCo ...............................................69
Figure 3-39 - Shaft-Hub Stress Profile for d=0.003’’, T=760K, ω=45,000rpm, Type A, ATi=AFeCo....................................................................71
Figure 3-40 – Pole-tip Radial displacement vs. RPM for T=760K, Type A, d=0.0015 and d=0.003’’’’, ATi=AfeCo .............................................71
Figure 3-41 – Shaft-Hub Radial displacement Profile for d=0.003’’, ω=45,000rpm, T=760K, t=2958hrs, Type A, ATi=AfeCo ...........................72
Figure 3-42 – Von Mises Strain Fringe Plot for d=0.003’’,ω=45,000rpm, T=760K, t=2958hrs, Type A, ATi=AFeCo ..........................................72
Figure 3-43 – Radial displacement vs. Time for d=0.003’’, T=760K, ω=45,000rpms, Type A, ATi=AFeCo ........................................................73
Figure 3-44 – Pole-tip Radial displacement vs. RPM for T=760K, Type A, d=0.0015 and d=0.003’’, ATi=AFeCo...............................................73
Figure 3-45 - Shaft-Hub Stress Profile for d=0.0015’’, T=780K, ω=35,000rpm, Type A, ATi=AfeCo...................................................................76
Figure 3-46 – Shaft-Hub Radial displacement Profile for d=0.0015’’, ω=35,000rpm, T=780K, t=943hrs, Type A, ATi=AFeCo...........................76
Figure 3-47 – Von Mises Strain Fringe Plot for d=0.0015’’,ω=35,000rpm, T=780K, t=943hrs, Type A, ATi=AFeCo ..........................................77
Figure 3-48 – Radial displacement vs. Time for d=0.0015’’, T=780K, ω=35,000rpms, Type A, ATi=AfeCo .......................................................77
Figure 3-49 – Pole-tip Radial displacement vs. RPM for T=780K, 760K, and 727K, Type A, d=0.0015’’, ATi=AfeCo ........................................78
Figure 3-50 - Shaft-Hub Stress Profile for d=0.003’’, T=780K, ω=35,000rpm, Type A, ATi=AfeCo.....................................................................78
Figure 3-51 – Pole-tip Radial displacement vs. RPM for T=780K, Type A, d=0.0015 and d=0.003’’’’, ATi=AfeCo .............................................80
Figure 3-52 – Shaft-Hub Radial displacement Profile for d=0.003’’, ω=35,000rpm, T=780K, t=642hrs, Type A, ATi=AFeCo.............................82
Figure 3-53 – Von Mises Strain Fringe Plot for d=0.003’’,ω=35,000rpm, T=780K, t=642hrs, Type A, ATi=AFeCo ............................................82
Figure 3-54 – Radial displacement vs. Time for d=0.003’’, T=780K, ω=35,000rpms, Type A, ATi=AFeCo ........................................................83
Figure 3-55 – Pole-tip Radial displacement vs. RPM for T=780K, Type AB1, ATi=AFeCo ..................................................................................83
Figure 3-56 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=588K, Tout=810K, ω=53,000rpm, Type A ................................................................85
Figure 3-57 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=588K, Tout=810K, ω=55,000rpm, Type A ................................................................85
Figure 3-58 – Radial displacement vs. Time for d=0.0015’’, Tin=588K, Tout=810K, ω=50,000rpms, Type A.....................................................86
Figure 3-59 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=588K, Tout=810K, ω=50,000rpm, Type A ................................................................86
Figure 3-60 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=727K, Tout=810K, ω=45,000rpm, Type A ................................................................89
Figure 3-61 – Radial displacement vs. Time for d=0.0015’’, Tin=727K, Tout=810K, ω=45,000rpms, Type A.....................................................89
Figure 3-62 – Pole-tip Radial displacement vs. RPM for Three Variable Temperature Profiles, Type A, d=0.0015’’.............................................91
Figure 3-63 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=810K, Tout=588K, ω=35,000rpm, Type A ................................................................93
Figure 3-64 – Radial displacement vs. Time for d=0.0015’’, Tin=810K, Tout=588K, ω=35,000rpms, Type A.....................................................93
Figure 3-65 – Shaft-Hub Radial displacement Profile for d=0.0015’’, ω=35,000rpm, Tin=810K, Tout=588K, t=2008hrs, Type A ......................94
Figure 3-66 – Two Lamina FE Mesh .......................................................................................................................................................................96
Figure 3-67 – Displacement vs. Time for 2 Lamina Model, Inner Radius unclamped, ω=45,000rpm, µ=0.............................................................99
Figure 3-68 – Deformation Plot at 5000 hrs for 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0 ............................................................100
Figure 3-69 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius unclamped, , ω=45,000rpm, µ=0 .....................................................101
Figure 3-70 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius unclamped, , ω=45,000rpm, µ=0 ..............................................................102
Figure 3-71 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius unclamped, , ω=45,000rpm, µ=0 ..............................................................102
Figure 3-72 – Displacement vs. Time 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0 ...............................................................................104

vii
Figure 3-73 – Deformation Plot at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0 ................................................................104
Figure 3-74 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius clamped, , ω=45,000rpm, µ=0 .........................................................106
Figure 3-75 - Von Misses Stress Fringe Plot 2 Lamina, Inner Radius clamped, , ω=45,000rpm, µ=0...................................................................107
Figure 3-76 - Von Misses Strain Fringe Plot 2 Lamina, Inner Radius clamped, , ω=45,000rpm, µ=0...................................................................107
Figure 3-77 – Displacement vs. Time for 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5 ...................................................................108
Figure 3-78 – Radial Displacement Profile for 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5...........................................................110
Figure 3-79 – Deformation Plot at 5000 hr for 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5 ...........................................................110
Figure 3-80– Radial and Tangential Stress Profile 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5 .....................................................111
Figure 3-81 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5 .............................................................112
Figure 3-82 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5 .............................................................112
Figure 3-83 - Displacement vs. Time for 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0.5........................................................................113
Figure 3-84 – Deformation Plot at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0.5 .............................................................114
Figure 3-85 – Radial and Tangential Stress Profile 2 Lamina, Inner Radisu clamped, ω=45,000rpm, µ=0.5 ........................................................115
Figure 3-86 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0.5 .................................................................117
Figure 3-87 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0.5 .................................................................117
Figure 3-88 – Displacement vs. Time for 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0 ......................................................................118
Figure 3-89 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.......................................................119
Figure 3-90 – Radial and tangential Stress Profile 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0 ........................................................120
Figure 3-91 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0 ................................................................122
Figure 3-92 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0 ................................................................122
Figure 3-93 – Displacement vs. Time for 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0..........................................................................123
Figure 3-94 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0...........................................................123
Figure 3-95 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0 ...........................................................124
Figure 3-96 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0 ....................................................................125
Figure 3-97 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0 ....................................................................125
Figure 3-98 – Displacement vs. Time for 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5 ...................................................................126
Figure 3-99 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5....................................................128
Figure 3-100 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5 ..................................................129
Figure 3-101 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5 ...........................................................130
Figure 3-102 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5 ...........................................................130
Figure 3-103 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5......................................................131
Figure 3-104 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5......................................................131
Figure 3-105 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5 ......................................................132
Figure 3-106 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5 ...............................................................134
Figure 3-107 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5 ...............................................................134
Figure 1.1 – Force and Boundary Conditions acting on the Shaft-Rotor System ...................................................................................................149
Figure 1.2 – Mesh Generation Variables................................................................................................................................................................149
Figure 1.3 – Variable Temperature Profile.............................................................................................................................................................150
Figure B-1 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=35,000rpm, Type A, ATi=AfeCo...................170
Figure B-2 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=40000rpm, Type A, ATi=AfeCo....................171
Figure B-3 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=45,000rpm, Type A, ATi=AFeCo..................172
Figure B-4 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=45,000rpm, Type A, ATi=AFeCo....................173
Figure B-5 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=50,000rpm, Type A, ATi=AFeCo....................174
Figure B-6 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=55,000rpm, Type A, ATi=AFeCo....................175

viii
Figure B-7 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=30000rpm, Type A, ATi=AFeCo...................176
Figure B-8 - Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=35,000rpm, Type A, ATi=AFeCo ..................177
Figure B-9 - Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=35,000rpm, Type A, ATi=AFeCo ....................178
Figure B-10 - Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=40000rpm, Type A, ATi=AFeCo ...................179
Figure B-11 - Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=25,000rpm, Type A, ATi=AFeCo ................180
Figure B-12 - Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=27000rpm, Type A, ATi=AFeCo .................181
Figure B-13 - Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=30000rpm, Type A, ATi=AFeCo .................182
Figure B-14 - Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=25,000rpm, Type A, ATi=AFeCo ..................183
Figure B-15 - Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=30000rpm, Type A, ATi=AFeCo ...................184
Figure B-16 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=588K, Tout=810K, ω=45,000rpm, Type A..185
Figure B-17 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=588K, Tout=810K, ω=47000rpm, Type A...186
Figure B-18 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=588K, Tout=810K, ω=48500rpm, Type A...187
Figure B-19 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=727K, Tout=810K, ω=30000rpm, Type A...188
Figure B-20 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=727K, Tout=810K, ω=35,000rpm, Type A..189
Figure B-21 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=727K, Tout=810K, ω=40000rpm, Type A...190
Figure B-22 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=810K, Tout=588K, ω=25,000rpm, Type A..191
Figure B-23 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=810K, Tout=588K, ω=27000rpm, Type A...192
Figure B-24 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=810K, Tout=588K, ω=30000rpm, Type A...193
Figure C-1 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=35,000rpm, Type AB1, ATi=AFeCo .............194
Figure C-2 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=40000rpm, Type AB1, ATi=AFeCo ..............195
Figure C-3 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=45,000rpm, Type AB1, ATi=AFeCo .............196
Figure C-4 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=50,000rpm, Type AB1, ATi=AFeCo .............197
Figure C-5 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=45,000rpm, Type AB1, ATi=AFeCo ...............198
Figure C-6 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=50,000rpm, Type AB1, ATi=AFeCo ...............199
Figure C-7 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=55,000rpm, Type AB1, ATi=AFeCo ...............200
Figure C-8 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=60000rpm, Type AB1, ATi=AFeCo ................201
Figure C-9 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=30,000rpm, Type AB1, ATi=AFeCo .............202
Figure C-10 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=35,000rpm, Type AB1, ATi=AFeCo ...........203
Figure C-11 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=40000rpm, Type AB1, ATi=AFeCo ............204
Figure C-12 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=35,000rpm, Type AB1, ATi=AFeCo .............205
Figure C-13 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=40,000rpm, Type AB1, ATi=AFeCo .............206
Figure C-14 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=45,000rpm, Type AB1, ATi=AFeCo .............207
Figure C-15 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=25,000rpm, Type AB1, ATi=AFeCo ...........208
Figure C-16 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=30,000rpm, Type AB1, ATi=AFeCo ...........209
Figure C-17 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=35,000rpm, Type AB1, ATi=AFeCo ...........210
Figure C-18 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=25,000rpm, Type AB1, ATi=AFeCo .............211
Figure C-19 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=30000rpm, Type AB1, ATi=AFeCo ..............212
Figure C-20 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=35,000rpm, Type AB1, ATi=AfeCo..............213

ix
TABLE OF TABLES

Table 1.2-1 Values of exponents in equation 1.2-6 for various physical mechanisms [8].........................................................................................7
Table 1.2-2 Approximate temperatures at which creep behavior becomes significant [14]. .....................................................................................8
Table 1.2-3 Creep Constants for HA50HS................................................................................................................................................................9

x
1 Introduction and Literature Review

1.1 Motivation

Using electric power to drive systems conventionally driven via mechanical, hydraulic, or
pneumatic power transfer systems is seen as the next technological advancement in aviation.
The more electric aircraft (MEA) will utilize electrical power in the form of power electronics,
fault tolerant electrical power distributors, and electronic flight control actuator systems to
generate a majority of the aircraft’s power [1,2].

Developing these advanced power technologies will significantly reduce the amount of
equipment and trained personnel required to maintain readiness for a given aircraft. These
advancements will be accomplished through the development of a magnetic bearing, an
integrated power unit (IPU), a starter/generator (IS/G) internal to an aircraft’s main propulsion
system, a fault tolerant electric Power Management and Distribution motor drive system and
electric primary flight control actuation. The IS/G’s benefits include elimination of various shafts
and gears in the engine along with the need for lubrication and related cooling. An increase in
specific power and range can be realized with certain aircraft configurations by reducing the
aircraft’s frontal area. A simplified schematic of the IPU is given in Figure 1-1. The IS/G is
placed aft of the compressor thereby adding work to the compressed airflow. The cooling air of
the compressor is actually quite hot, some estimates of its temperature are around 600°F. The
spinning of the rotor, as well as magnetic core losses, increase the rotor’s operating temperature.
These advancements would also eliminate the need for conventional components such as oil
lubrication/ball bearing systems, turbine inlet stator/nozzles, instrumentation, accessory gearbox
driven fuel and lubrication pumps, and hydromechanical fuel controls [1,3-5].

There are many technological challenges to overcome in designing an IPU and an IS/G. There
are unconfined magnetic fields and stray electronic currents when dealing with rotating
machinery [6]. There is the possibility of large lateral forces directed upon the central shaft [7].
The IPU and the IS/G require a high power density in order to produce sufficient power to run

1
Figure 1-1 – Integrated Power Unit Schematic

2
the secondary power functions of an aircraft. The rotors and stators are turbomachines
comprised of a stack of soft magnetic thin laminas made of an iron-cobalt alloy. A typical IPU
configuration might include six stator poles and four rotor poles as shown in Figure 1-2. The
required operating speeds and temperatures dictate that a high strength magnetic material will be
required in order to maintain integrity.

The first step towards implementing an IPU design utilizing soft magnetic materials is to gain an
understanding of the mechanical behavior (i.e. creep) of the material. This was accomplished by
Fingers [3]. The next step, and the focus of this work, is to create a finite element (FE) tool
capable of modeling the stress and deformation fields resulting from long term creep. This tool
can be used to predict the effects of geometry and operating environment on the deformation of
the rotor stack. The final step in developing a proven design tool (not covered in this thesis)
involves laboratory tests of various rotor configurations. These tests could be used to refine the
FE tool, and thus develop a design tool capable of evaluating various rotor configurations and
predicting their real world usefulness.

1.2 Creep Phenomena

Elastic and plastic deformations are usually idealized as appearing immediately upon loading.
Creep is a time dependent phenomenon that occurs from the application of a constant load over
time. For metals and crystalline ceramics, creep is important in a temperature range generally
between 30 and 60% of the material’s absolute melting temperature [8]. “Creep strain” is
defined as the strain that occurs after the initial elastic and, in some cases, plastic deformation.

The typical strain versus time behavior is shown in Figure 1-3. The first, or primary, stage
contains the instantaneous elastic and plastic strain followed by the increasing addition of creep

strain. The primary stage, or strain hardening, is defined by the initial large strain rate, ε! = .
dt
The amount of strain hardening depends on the number and types of dislocations. The initial

3
Figure 1-2 – Typical Rotor/Stator Configuration [5]

Figure 1-3 – Typical Creep Strain vs. Time Curve

4
strain hardening is opposed by an increased flow due to softening from recovery and
recrystallization. The strain rate decreases and typically approaches a constant value, where
strain hardening is balanced by the increased flow. This is characteristic of the secondary, or
steady state, stage. Towards the end of the secondary stage ε! increases unstably as the specimen
approaches failure; this is the tertiary stage. The final stage ends in rupture of the body; it is
associated with the growth of microcracks and/or necking in the material [9].

Andrade [10,11] first studied high temperature creep in the early part of the 20th century.
Andrade’s studies revealed that in the primary creep region, the creep rate depends on time,
especially at high temperatures and stresses. He proved that
ε = ε0 + β ⋅tm (1.2-1)

where ε0 represents the instantaneous strain, t is the time, and β is a material constant; the
Andrade creep is defined as m equal to 1/3.

Many physical mechanisms affect creep rate for a given combination of stresses and temperature.
One of the first mechanisms analyzed was diffusional creep, which is typically active at very low
stresses and very high temperatures. Diffusional creep is characterized by grain shape changes.
These changes are due to atoms being transported from boundaries experiencing compressive
normal stress to those experiencing tensile normal stress, and are accommodated by grain
boundary sliding [12,13]. The typical primary creep curve, exhibited in Figure 1-3, depicts a
high creep rate. This type of creep can be separated into low temperature (logarithmic) creep
and high temperature transient creep. The general equation describing the logarithmic creep is
ε = ε 0 + ε 1 [ln (t + t 0 )]
s
(1.2-2)

where ε is strain, and ε 0 , ε 1 ,s, and t0 are constants. This form was derived from the theory of
exhaustion creep presented in [14]. A second model for transient creep proposed by Orowan
assumes that the creep rate is proportional to the stress, and is represented by
σ 
ε =   − e ch (t0 −t ) (1.2-3)
h
where σ is the stress, h is the coefficient of work hardening and c is a constant. Assuming that ε
equals zero at time equal to zero it can be shown that

5
ε! = cσe − cht (1.2-4)

This equation indicates that the strain rate decays exponentially with time rather than the
observed logarithmic decay. The logarithmic decay is predicted if it is assumed that the creep is
thermally activated and that the activation energy is linearly dependent on the stress [14].

In addition to the two primary creep models presented here, Nabarro [14] has suggested two
other mechanisms that effect primary creep in polycrystals. The first is grain-boundary sliding
which can produce strains on the order of elastic strains. The second is based on the Schmidt
factor which is defined as the “ratio of the maximum resolved shear stress on a glide system in a
crystal to the applied tensile stress” [14]. This value can vary between 0.275 and 0.5, depending
on the crystal orientation. It is believed that the first mechanism could not lead to large
differences in primary creep since it produces strains smaller than the observed primary creep
strains. The second mechanism, on the other hand, could well have a substantial effect on the
primary creep. However, it is difficult to prove. An examination of typical creep curves for
aluminum or steel indicate primary creep extensions of 10 to 20%, while for single-crystal
superalloys the creep extensions are around 0.2%. These differences may be attributed to
material properties and loading conditions [14].

Dowling [8] uses rheological models to describe a relationship between the stress, the strain and
the time:

  
 − E2t 
 σ  σ ⋅t   σ 
ε =   +   +   1 − e  η2   (1.2-5)
η  
 1   1   E2
E  

σ  σ ⋅t 
where   is the instantaneous elastic strain due to loading,   is the steady state strain,
 E1   η1 

  
 − E2t 
σ 
and   1 − e  η2   is the transient strain. This equation describes diffusional dominated
 
 E2  
creep, as proposed by Nabarro [14].

6
The majority of the creep strain occurs while the body deforms at a constant strain rate.
Therefore, this can be considered a major factor where creep is concerned. This stage is
generally considered to be thermally activated. The following equation represents a general
power law relation for steady-state strain rate in metals and crystalline materials:
 −Q 
 Aσ m   RT 
ε! =  q e (1.2-6)
d T 
where σ represents the stress, d the average grain diameter, T is the absolute temperature, and R
is the universal gas constant. The stress coefficient A, exponents m and q, and activation energy
Q depend on the material. Typical values of m and q are listed in Table1.2-1. Much of the work
on creep has been focused on the steady-state creep mechanisms [15-19]. The temperature
dependency is obvious in equation 1.2-6. It is commonly observed that as the temperature of a
material increases, its strength decreases. The melting temperature thus has a substantial
influence on the creep behavior of a material. It is indicated in [14] that the homologous
temperature, the ratio of the absolute temperature (Ta) to the absolute melting temperature (Tm),
is useful for comparing the creep characteristics of various materials. Significant creep is
generally experienced at temperatures of 0.5 Tm. This is typically referred to as the Andrade
creep or the high-temperature creep. Table 1.2-2 lists the temperatures where creep becomes
significant for various materials.

Table 1.2-1 Values of exponents in equation 1.2-6 for various physical mechanisms [8].
Name of Mechanism m q Description

Diffusional flow (Nabarro-Herring Creep) 1 2 Vacancy diffusion through the crystal lattice

Diffusional flow (Coble creep) 1 3 Vacancy diffusion along grain boundary

Grain boundary sliding 2 2 or 3 Sliding accommodated by vacancy diffusion


through the crystal lattice (q = 2) or along the
boundaries (q = 3)

Dislocation creep (Power law) 3 to 8 0 Dislocation motion

This study involves a material subjected to temperatures above 0.5 Tm and moderate to high
stresses. It has been suggested that the creep occurring in this regime is dictated by edge

7
dislocation climb away from dislocation barriers [20]. Fingers [3] found that the following
formula fitted well to his test data:
 Aσ n  −RTQ
ε! =  e (1.2-7)
 T 
where A is a material constant, Q is the activation energy associated with lattice diffusion, and n
is the stress exponent. These material properties are presented in Table 1.2-3 for the iron-cobalt
alloy HA50HS which is the focus of this study. The above equation agrees with [19] that
indicates the following relation can be used to represent creep for high temperature and high
stress conditions:
Q

ε! = Se RT
ϕ (σ ) (1.2-8)
In this study it is assumed that
ε total = ε e + ε Tr + ε TH + ε c (1.2-9)

where εtotal is the total strain, εe is the elastic strain, ε Tr is the transient strain, ε TH is the strain

resulting from thermal expansion, and εc is the steady state creep strain. This study only
examines the elastic and the steady state creep strain.

Table 1.2-2 Approximate temperatures at which creep behavior becomes significant [14].
Material Type °F T/Tm

Aluminum Alloys 400 0.54

Titanium Alloys 600 0.30

Low Alloy Steels 700 0.36

Nickel- and cobalt-base superalloys 1200 0.56

8
Table 1.2-3 Creep Constants for HA50HS
Creep Type Q(cal/mol) n A (K/hr (psi)^n)

A 98,851 3.83 4.940E+07

B1 99,318 8.49 1.060E-15

B2 87,259 13.54 1.020E-42

C 72,596 25.33 1.560E-105

AB1 95,329 4.02 7.210E+05

1.3 Finite Element Method

The concept of the finite element analysis (FEA) has been in use in some form for centuries.
The simple idea is to replace a complex problem with a simpler one that represents the true
solution to a desirable degree of accuracy. It is apparent that ancient mathematicians employed
finite element (FE) like problem solving techniques to many physical problems. With the advent
of the computer FEA can now be applied to more broad and sophisticated problems.

The most obvious indication of ancient mathematicians utilizing FE techniques are found in
early attempts to calculate π. Chinese literature written in the beginning of the first century A.D.
indicates that the Chinese were aware of certain geometric theorems about the same time as the
Greeks. These theorems led Tsu Chung Chilk, a Chinese engineer, to estimate π to be between
3.1415926 and 3.1415927 by 480 A.D. [21]. This was most likely accomplished by representing
the area of a circle with inscribed and circumscribed rectangles, or finite elements (as seen latter
in this section). Archimedes used the finite element technique to estimate the volume of various
solids. These are just a few of the more obvious early uses of the finite element technique.
Today FEA is used to solve more complex problems.

Martin [21] outlines the development of the modern finite element method (FEM) by following
the evolution of aircraft. Aircraft design and optimization has always presented engineers with
some of their most difficult problems. The need for lightweight, high performance, and safe
planes initiated the development of refined structural analysis techniques. New analytical

9
methods for analyzing static and dynamic problems associated with aircraft became available by
the 1940’s. This development coincided with the development of high strength, lightweight
alloys used in early airplanes. By the late 1940’s jet power aircraft began to appear. The advent
of the jet led to a whole new set of problems in the design of aircraft. It became apparent that the
previous techniques were no longer adequate to account for the increased speed and design
changes (swept wings, the delta wing) that went with higher speeds.

The mathematical formulation of physical problems is often difficult. The resulting models may
include differential equations and require an understanding of the physical process involved. In
general, derivation of governing equations is not difficult, though obtaining closed form
solutions can often be a formidable task. Approximate methods are crucial to solving problems
where an exact solution is not possible. Methods frequently used in literature include the finite
difference method and variational methods such as the Galerkin method.

The finite difference approximation of a differential equation involves expressing derivatives in


terms of difference formulae obtained by the Taylor series expansion of a function. This requires
values of the unknown variable at discrete points in the domain. Boundary conditions are
imposed on the resulting algebraic equations. Thus the governing equations are satisfied at the
discrete points [22].

Variational methods place differential equations into an equivalent weighted-integral form. The
solution is assumed to be a linear combination of simple functions and undetermined
coefficients. These coefficients are determined by satisfying the weighted-integral equations.
The various methods differ in the selection of the basis functions. One problem with variational
methods lies in the difficulties inherent in constructing basis functions for arbitrary domains.

To better understand the basic concepts of the finite element method, Reddy [22] considers the
problem of estimating the circumference of a circle. One way to obtain an estimate of the
circle’s circumference is to approximate it by a sum of easily measurable line segments. This
simplistic approach can be used to outline some of the basic concepts and terminology involved
with finite element analysis. First, the domain is separated into a finite, n, number of

10
subdomains. This process is referred to as the discretization of the domain. Each of the
subdomains is called an element. The elements are connected to one another via points called
nodes. The collection of all of the elements and nodes is referred to as the finite element mesh.
If all of the elements are (as in this case) of equal length the mesh is referred to as a uniform
mesh otherwise it is called a non-uniform mesh, see Figure 1-4.

The next step involves determining the element equations. The required properties, in this case
the element length, are computed for each element (Ωe) individually. If he represents the length
of a typical element (see Figure 1-4), then:
1 
he = 2 R sin  θ e  (1.3)
2 
where R is the radius and θe < π is the angle subtended by the element.

The third phase involves assembling the element equation to find a solution. The approximate
solution is obtained by combining the element properties. In this case the solution is obtained by
realizing that the perimeter of the polygon, is simply the sum of the individual element lengths.

Pn = ∑e =1 he
n
(1.3-5)


Pn equals the approximate circumference. For a uniform mesh, where θ e = , then:
n
  π 
Pn = n 2 R sin    (1.3-6)
  n 
The final steps are convergence and error estimation. Since this problem has a closed form
solution, p = 2πR, the error can be easily calculated. It is relatively simple to show that Pn will
converge to the exact solution as n→∞. The error is simply the difference between the length of
the circular segment (Se) enclosed by the element and the element length.
E e = S e − he (1.3-7)

Therefore, the estimated error for an element and the total, or global, error is:
 2π  π 
Ee = R  − 2 sin   (1.3-8)
 n  n 

11
Figure 1-4 – Estimation of the Circumference of a Circle

12
E = 2πR − Pn (1.3-9)
The circumference of the circle can be estimated as closely as one wishes by using the
summation of a finite number of piece-wise linear functions.

The simple problem of estimating the circumference of a circle brings to light some of the
advantages of the finite element method. This method allows one to solve structural analysis and
other problems for various materials or complex domains.

Other features of the finite element method are not readily apparent in the above example. Nath
[24] summarizes some of the main attributes of the FEM. The flexibility of the size and shape of
the elements allows for modeling of complex structures, such as bodies with one or more holes.
Domains with variable material or geometric properties are simple to handle. Even non-linear
and time dependent properties can be accommodated.

There are three general sources of error inherent in the finite element method. The first source,
which is the only one apparent in the example, comes from the approximation of the domain.
The second source comes from the approximation of the unknown function in terms of a finite
set of basis functions. The third source of error comes from the necessity of utilizing computers
for complex finite element problems. Numeric integration and round off cause this error. For
nonlinear problems, it is quite difficult to estimate these errors, though for relatively simple
problems (like the example) it can be done.

When the finite element method is used to analyze a time dependent problem, such as creep, the
problem formulation is performed in two stages. The first stage involves the approximation of
the partial differential equations as a set of ordinary differential equations in time. In the second
stage the ordinary differential equations are solved or approximated by algebraic equations
which are solved at discrete values of the time t

In summary, the finite element method provides a powerful numeric technique for solving
boundary-value problems over a complex domain. The finite element method has been used to
analyze the mechanical deformations of a rotor lamina comprised of a soft-magnetic material and

13
subjected to a high temperature and high stress operating environment. The results of this
analysis are presented in this report.

1.4 Previous Work

This section briefly covers the work performed towards developing the more electric aircraft
(MEA). An overview of the United States Air Force’s initiative, an assessment of relevant
laboratory tests, and the use of FEA in investigating creep problems are included. There are
many technological improvements being pursued for implementation in the MEA. This report
focuses on the development of the IPU and IS/G, specifically the shaft-rotor system.

In the early 1940’s engineers at Patterson Field determined that, due to a lack of electrical power
generation capability, hydraulic power was the only feasible method to provide aircraft
secondary power. To this day commercial and military aircraft employ this and various other
means of power generation. These systems have improved in performance and reliability but
still account for a significant amount of maintenance and failures. As technological capabilities
grew, the Air Force started a program to research and demonstrate the MEA. Studies show that a
MEA would experience appreciable improvements in maintainability, reliability, supportability,
cost and weight. As the initiative continued it was realized that other vehicles such as naval
vessels, and various Army vehicles, would benefit from the conversion to electric power
generation [25].

The program was structured as a three generation time and technology availability based
research and development project. The first phase (Generation 1) was completed in 1998. This
phase established the electrical needs necessary for the removal of aircraft hydraulics.
Generation 2 projects aim to develop significant advancements in electrical power generation
capabilities by 2005. This phase involves demonstration of the MEA. Improvements of around
1400% in reliability and 200% in power density over conventional aircraft are expected. Finally,
generation 3 projects aim to power all non-propulsive aircraft functions via the IS/G and IPU,
driven electrically, by 2012. The high power levels needed to accomplish these tasks will be
possible through the use of electric generators using superconducting alloys [25,4].

14
A series of papers by Wahl [26-30] outlines an analysis that combines analytical models with
laboratory test data to predict deformations of a spinning rotor subject to high temperatures and
stresses. Wahl made the following assumptions:

1. The elastic deformations are negligible as compared to the creep deformations.


2. A steady-state condition of stress is reached. Therefore, the initial transient condition during
the start-up is neglected.
3. A biaxial state of stress (plane stress) comprised of radial stress (σr) and tangential stress (σθ)
exists.

4. Creep rate is given by ε! = F (σ ) f (t ) = kσ f (t ) .


n

Assumption 1 is valid for a spinning disk. Assumption 2 indicates that the consequences of
starting and stopping a rotor are negligible. Start-up, the time required to accelerate to full
power, is around 2-seconds [3]. Compared to an operational lifetime of 5,000 hours, neglecting
the creep strains during the start-up period should have little impact. The third assumption is
valid for thin laminas. If the whole rotor is modeled as a solid body, a plane strain state of
deformation would be more appropriate. The last assumption appears to be validated by the test
data given in [3].

Expressions for different components of the strain rate tensors are [26]:
 1 
 σθ − σ r 
ε!θ = ks n f (t ) 2

 s 
 

 1 
 σ r − σθ 
ε! r = ks n f (t ) 2
 1.4-1
 s 
 

 − (σ r + σ θ )
ε! z = ks n f (t ) 
 2s 

15
where s is the equivalent stress which is dependent on the yield criterion. The yield criteria
examined in [28,29] are the von Mises yield strength criterion, the Tresca yield strength
criterion, and a combination of the two. Wahl concluded that the von Mises criterion yielded
predictions that were too conservative and that the Tresca criterion best represented the results.
A separate study [31] involving creep in a class of shell closures also found the Mises criterion
and its associated flow rule to give low (unsafe) calculated values of creep deformations. This
conclusion is also supported by D’Isa’s [32] study of creep in rotating disks comprised of
chemically pure lead. The equivalent stress for the Tresca criterion is:
s = σθ 1.4-2
Strain rate relations (1.4-1) can then be written in the following form:

u!  σ 
ε!θ = = kσ θn 1 − r  f (t )
r  2σ θ 

du! σ 1
ε! r = = kσ θn  r −  f (t ) 1.4-3
dr σθ 2 
ε! z = −(ε!θ + ε! r )
If the flow rule associated with the Tresca criterion is assumed these equations become:

u!  σ 
ε!θ = = kσ θn 1 − r  f (t )
r  2σ θ 
ε! r = 0 1.4-4
ε! z = −ε!θ

Wahl has shown that ε! r = 0 is correct except at the inner and the outer diameter. In [27] a disk
with a central hole was subjected to a load resulting in a stress σro on the outer radius of the disk.
The resulting equations for the radial and tangential stresses are
 r −r  −n1
σ θ = ασ θAVG  αo iα r
 1.4-5
 ro − ri 

σ θAVG (ro − ri )  r α − riα  ρω 2 3


σr =  α α 
− (r − ri3 ) 1.4-6
r  ro − ri  3r
n −1
where α = , and
n

16
ro σ ro 1  r 3 − ri3 
σ θAVG = + ρω 2  o 
 1.4-7
ro − ri 3  ro − ri 
These equations assume that the loads caused by the rotor poles are spread uniformly along the
outer radius of the hub. This obviously is a simplification since the stresses would exhibit
dependence on θ as a result of loads acting on the pole. The FE models presented in this report
show this phenomena. Experimental results from [3] indicate that, for the iron-cobalt alloy
studied in this report, k (from equation 1.4-4) can be represented as:
Q

Ae RT
k= 1.4-8
T
where A is the stress coefficient, Q is the activation energy, R is the universal gas constant, and
T is the temperature. For a steady-state condition, where f(t) = 1, the tangential strain can be
represented as:
 −Q 
 σ  Ae RT  n
ε θ = 1 − r  σ θ t 1.4-9
 2σ θ  T 
 
In [28] Wahl examines a variety of geometries and loading conditions. These include a radial
variation in thickness, varying the ratio of the inner and the outer diameter, and a radial variation
in temperature. The latter models temperature as a function of the radius such that the creep rate
at the outer diameter is ten times that at the inner diameter. A similar case is examined in this
report except that the temperature variation is considered a linear function of the radius. Wahl’s
study indicates that modeling temperature as a function of the radius results in a greater non-
uniformity in the stress distribution as compared to that in a constant temperature disk.

The effects of the initial transient period are studied in [30]. No appreciable difference in the
total creep deformation was noted as a result of accounting for the transient period. It was
concluded that for cases where creep strains are significantly greater than the elastic strains it is
not necessary to include the effect of the transient period. The transient period would only be of
consequence when the creep strains were on the order of the elastic strains. This report shows
that when the shaft-rotor interference fit is added to the model, the “operational” creep strains1

1
“operational” creep strains refer to the creep strains noted prior to failure.

17
are indeed on the order of elastic strains. This leads to the question as to whether the creep
strains during the transient period can be neglected. This is discussed in detail latter in the
report.

Wahl’s work [26,27] combined laboratory tests with an attempt to derive analytical
representations of creep deformations in rotating disks and rotors. Mendelson [33,34] took the
next step and used a numerical method to calculate the creep deformations. A method of
successive approximations was used. It begins with the elastic solution followed by calculating a
first approximation of the plastic strains. These strains are used to estimate the stresses and total
strain distribution. Finally, the plastic strains are recalculated and the process is repeated to
convergence [33]. This approach produced calculated stresses and strains in good agreement
with the data presented in [26].

Yang [35] approached the problem of estimating creep deformations for a rotating disk by using
an incremental FE approach. This involved the theory of viscoplasticity to formulate the
constitutive equations. The principle of virtual work is used to derive a viscoplastic pseudo force
and a Norton type time hardening intrinsic time scale is used for the viscoelastic creep rule [35].
The results of this analysis are in good agreement both with the data obtained in [26] and the
numerical results in [33]. This paper clearly indicates the usefulness of using the FEM to predict
creep deformations of a spinning rotor in a high temperature and a high stress environment.

This report uses the FEA program ABAQUS to examine the use of iron-cobalt alloys in rotor
construction.

1.5 Material

According to Fingers [3] iron-cobalt alloys are the only materials that should be used for high
stress and high temperature applications. Cobalt is the only alloying material that significantly
increases the Curie temperature as well as the saturation magnetization of iron [36]. A good
candidate must also posses a high permeability, as well as a high Curie temperature and

18
saturation magnetization. The initial and maximum permeabilities are two values used to
categorize soft magnetic materials. An iron alloy with approximately 50% cobalt has a high
magnetization as well as high permeability as compared to alloys with lower percentages of
cobalt [3].

The mechanical and electrical properties must also be considered when designing a switched
reluctance starter/generator rotor and stator. The geometry affects the mechanical behavior and
windage losses. Hysteretic loss is a material property and the eddy current loss is proportional to
the lamina thickness and square of the frequency and inversely proportional to the material’s
resistivity. These are both components of the core losses, which are associated with a particular
design [3]. To prevent high losses the rotor core laminas must be insulated from each other.
This is sometimes accomplished via an organic varnish [37]. The most common procedure is to
anneal in an oxidizing environment, such as air. Fingers obtained the material properties for an
iron-cobalt alloy that was annealed in an oxidation free, dry hydrogen environment. The
material was then oxidized in a cooler air environment [3]. The oxidation results in deep blue
iron oxide approximately 50 nanometers thick. For more complete details of treating the
material, the reader is referred to publications[3,38].

The material properties used in this study are for Hiperco Alloy 50HS (HA50HS). The
melting temperature is advertised to be 1500°C and the Curie temperature is 940°C. The
saturation induction is 23,500 Gauss at an applied field of 200 Oesterds [3,39]. The family of
Hiperco Alloys (HA) has been used in a multitude of applications based on the mechanical and
magnetic requirements. HA50 has typically been used in motor and generator applications
where core losses are sometimes compromised in order to obtain improvements in mechanical
behavior [3]. When core loss is the key design factor, HA50A is used.

Studies by Carpenter Technology Corporation indicate that grain size is the main variable
affecting material properties. HA50 contains niobium, which has growth-impeding properties,
and exhibits reduced grain growth rates [3]. These smaller grain sizes result in a higher strength
material. HA50A has larger grain sizes than HA50 and therefore, exhibits a better magnetic

19
performance. HA50HS contains the highest addition of niobium to retard grain growth during
material preparation, as can be seen in Table 1.6-1.

Table 1.6-1 Chemical Composition (wt%) of Carpenter’s Fe-Co Alloys [3].


Hiperco Co Fe V Nb Si Mn C
Alloy

HA50 48.71 Balance 1.93 0.05 0.03 0.02 0.007

HA50A 48.85 Balance 2.03 <0.002 0.03 0.06 0.003

HA50HS 48.75 Balance 1.90 0.30 0.05 0.05 0.01

1.6 Problem Statement

The development of a MEA requires the use of advanced soft magnetic materials capable of
performing in high temperature and high stress environments. These materials will be used in
the construction of rotors and stators for use in IPU and switched reluctance IS/G. There is now
a need for a design tool to analyze potential rotor configurations for use in these machines. A
first step towards this goal is the design of a FE tool that is capable of receiving input such as
rotor geometry, material properties, temperature environment, and rotating speeds. This tool
would then create a finite element model (FEM) which can generate theoretical stress and strain
fields as well as creep deformations. The results can be used to investigate the effects of
geometry, material properties and operating environment on the integrity of the rotor.

1.7 Approach
We will attempt to add to the body of work being done to analyze creep deformations of soft
magnetic iron-cobalt alloys. This will be accomplished by the development of a code to generate
ABAQUS input files for the analysis of single lamina, and the analysis of the effects of friction
and temperature variations in laminated rotors.

20
Generation Code
A code has been developed that allows the user to input variables such as rotor geometry,
material properties and operating environment conditions. Using this input the code will
generate a complete ABAQUS input file (which includes the FE mesh, loads, material
definitions and output formats). These files will be utilized by ABAQUS to generate stress and
deformation profiles for various rotor configurations.

ABAQUS
ABAQUS is a finite element code developed and distributed by Hibbet, Karlsson & Sorensen,
Inc. This study utilizes the isotropic elastic and the power law creep material models to perform
the analysis.

Post Processing Code


A code has been developed to assist in the analysis of the results obtained from ABAQUS. This
code generates files for use in various programs to compile the data into useful graphs and visual
aids. Some of the files are used to generate stress, strain and radial displacement time histories
and profiles. Database programs with plotting capabilities, such as Microsoft Excel can be used
with these files. Other files are used to generate 3-D deformation plots as well as stress fringe
plots. PATRAN software could be used in conjunction with these files.

Single Lamina Models


The first phase of this project involves the examination of one-lamina thick shaft-rotor models.
A comprehensive examination of the effects of temperature, time, material properties and
rotational speed will be performed. Initial models will assume that the temperature is constant
and independent of the radial position. Then a radial variation in temperature will be examined
by assuming that the model has reached a steady-state condition; the radial variation in
temperature can then be considered by modeling successive radial layers using temperature
dependent material properties.

21
Multiple-Laminae Models
The final stage of this project will be an attempt to model the interaction between two or more
laminae. This will be accomplished by assuming Coulomb friction between the laminae. This
assumes that there is no slipping between the contacting surfaces and that more complex models
of friction, such as kinetic friction, are not required. This is a first attempt to understand the
effects of frictional forces.

22
2 Analysis Procedure

2.1 Mesh Generation and Load Calculations

The first step in the analysis is the generation of the FE mesh. The author developed a
FORTRAN code that generates this mesh from data in a user-input file. Explicit details of this
code (genmesh.F) and required user-input file can be found in the user’s manual included in
appendix A. Genmesh.F is capable of generating the FE mesh including various load conditions
as a complete input file for ABAQUS.

As stated in the Introduction and Literature Review section, the FE mesh is dependent on the
geometry of the domain and type of element utilized. A typical eight noded brick element is
used by genmesh.F to represent the geometry of the rotor, see Figure 2-1. The geometry is
defined via a few simple dimensions input by the user. This code is capable of generating FE
meshes for a shaft-rotor system (with or without the shaft and rotor poles). The required inputs
to generate these meshes are (see Figure 2-2): the radius of the hole (r1), the radius of the hub
(r2), the radius of the pole (r3), the shaft wall thickness (tbt), the thickness of the lamina (t) and
the width of the poles (tpl). These variables define the boundaries of the domain; now the
number of elements, and the number of nodes, need to be defined. Figure 2-3 depicts how the
user defines the number of elements. The variables NRADSEG and NTHSEG define the
number of elements in the radial and the circumferential directions respectively. Note that the
number of elements in the radial direction (NRADSEG) represents the number of elements,
along a radial line, from the central hole (r1) to the hub (r2). Note also that the number of
elements in the circumferential direction must be evenly divisible by four. The last two variables
used to generate the FE mesh on the rotor lamina are NTHPL, the number of elements in the
width (tpl) of the rotor pole and NTBSEG, the number of elements in the radial direction of the
shaft.

In the remainder of this report the shaft-rotor system will be referred to in four ways: 1. the shaft,
2. the hub, which is the portion of the rotor not including the pole-tips 3. the poles and 4. the
rotor, which is the combination of 2 and 3.

23
Figure 2-1 – Typical Eight Noded Finite Element

Figure 2-2 – Shaft-Rotor Geometry

24
Figure 2-3 – Shaft-Rotor Mesh Generation Variables

25
With the FE mesh created, the next step is to calculate the forces acting on the lamina, see Figure
2-4. The main load applied to the rotor in this study is the load generated by rotation. Other
loads include those incurred from the interference fit and the clamping force applied to the rotor
stack, see Figure 2-4.

The centrifugal force acting on an element of a spinning disk can be represented as follows:
(
dFel = ρrω 2 (tdA) ) (2.1-3)

where t is the thickness of the element, ρ is the mass density, ω is the rotation speed, r is the
radial position of the element centroid, and dA represents the infinitesimal area of the element.
In order to use these forces in ABAQUS, they need to be separated into their x and y coordinates.
( )
dFelx = ρrω 2 (tdA) cos θ (2.1-4)

dFely = (ρrω )(tdA)sin θ


2
(2.1-5)

Transferring these to nodes on the element, integrating and replacing (rcosθ) with x, (rsinθ) with
y and dA with dxdy, we obtain
(
Felxa = ρω 2 t )∫ ∫ ψ a xdxdy (2.1-6)

(
Felya = ρω 2 t )∫ ∫ ψ a ydxdy (2.1-7)

where the superscript a denotes the node number at which the force is applied. The next step is
to transform the integration from the local element onto a master element, see Figure 2-6. In
order to do this we use the following mapping functions:

x = ∑α =1 xα ψ α (s, t )
nodes
(2.1-8)

y = ∑α =1 yαψ α (s, t )
nodes
(2.1-9)

dxdy = Jdsdt (2.1-10)


where J is the Jacobian of the transformation, and the master element coordinate system is
represented by s and t. Using these relations the x and y components of the centrifugal forces
become:

( )∫ ∫ ψ ∑ x ψ (s, t )Jdsdt
1 1
Felxa = ρω 2 t a α α (2.1-11)
−1 −1 α

26
Figure 2-4 – Forces and Boundary Conditions Acting on the Shaft-Rotor System

Figure 2-5 - Centrifugal Force acting on a Single Element

27
Figure 2-6 - Node Numbers and interpolation functions for a Lagrange rectangular element

28
( )∫ ∫ ψ ∑ y ψ (s, t )Jdsdt
1 1
Felya = ρω 2 t a α α (2.1-12)
−1 −1 α

These relationships can then be simplified to the following form by using quadrature formulae.

( )∑ w
n int
Felxa = ρω 2 t int f inta x , a = 1, nodes (2.1-13)
int =1

( )∑ w
n int
Felya = ρω 2 t int f inta y , a = 1, nodes (2.1-14)
int =1

where nint is the number of integration points, f inta x and f inta y correspond to ψ a ∑ xαψ α (s, t )J
α

and ψ a ∑ yαψ α (s, t )J , and wint is the weight corresponding to the integration (gauss) point int.
α

Four integration points are used in each direction as shown in Figure 2-7. Once the nodal loads
are calculated for each element individually they must be assembled into a global load vector.
This is accomplished by accounting for nodes common to multiple elements and by summing the
nodal loads from every relevant element, thereby generating nodal loads for the FE mesh. The
analysis was performed on one surface of the element while the model being developed requires
both the top and bottom surfaces. However, the loads on the top and bottom surface of the
lamina are equal.

The rotor stack is mounted onto the shaft using an interference fit. This is accomplished by
forcing the cylinders together with a press or by shrink fitting [40]. Shrinking is accomplished
by heating the rotor stack or cooling the shaft in order to allow the shaft to slide inside the rotor
hole. Both components are then allowed to cool/heat to ambient temperature resulting in the
interference fit. This method induces a stress field in the shaft-rotor system. The stress field can
be calculated analytically [40]. These equations are derived for the interference fit between two
equal length cylinders. This allows one to calculate the interference fit based on material
properties and the contact pressure between the cylinders.

The approach is to model the interference fit as a combination of an externally loaded and an
internally loaded cylinder, see Figure 2-8. The equation for determining the interference fit in
terms of the contact pressure is:

29
Figure 2-7 - Integration (Gauss) Points for a Quadrilateral Element

30
Figure 2-8 Radial and Tangential Stress Profiles for Internally and Externally Pressurized Cylinders

31
(
 r  r 2 + r 2  r)
d = Ps  2  32 22 + µ 1  −  2
(
 r22 + r12
 2 −
)
µ


( ) ( )
 r −r2  (2.1-15)
 E1  r3 − r1
2
  E2  2 1 
where d represents the interference fit or the sum of the decrease in the outer radius of the inner
cylinder and the increase in the inner radius of the outer cylinder as a result of the fitting process.
E1 and E2 represent the elastic moduli of the materials of the inner and outer cylinders
respectively, and µ1 and µ2 represent Poisson’s ratio for the two cylinders. Ps is the contact
pressure. This equation can be solved for the contact pressure.
d
Ps = (2.1-16)
(
 r2  r + r2 2
)+ µ   r2 (
 r22 + r12) 
   
1 −
 2 − µ 
2 
3 2

(
 E1  r − r
3
2
1
2
)   E2 (
 r −r2
 2 1 ) 
Equations for non-rotating cylinders subjected to internal and external pressure loads can be used
to represent the radial and tangential stresses, see Figure 2-8. For an internally loaded cylinder
the following equations apply:
K 1 ro2
σ r = K1 − (2.1-17)
r2
K 1 ro2
σ θ = K1 + (2.1-18)
r2
Ps ri 2
where K1 equals . The subscripts i and o indicate respectively the inner and outer radius
ro2 − ri 2
of the cylinder. Equations for the externally loaded cylinder are:
K 1 ri 2
σ r = K1 − (2.1-19)
r2
K 1 ri 2
σ θ = K1 + (2.1-20)
r2
− Ps ro2
where K1 equals .
ro2 − ri 2

These relations are for two cylinders and do not account for the effects of the poles of a rotor. It
is intuitive that rotor poles induce a dependence on theta in the stress fields. An examination of
the stress fields induced for the internally loaded cylinder clearly shows that the stresses near the

32
outer radius are small compared to the stresses at the inner radius. Therefore, one can assume
that the theta variance will be only found near the outer radius and consequently negligible. This
assumption is further supported by realizing that the loads incurred from the high rotation speeds
will further overwhelm any circumferential variance occurring as a result of the interference fit.
Based upon this assumption, the effects of the poles can be accounted for by smoothing the area
of the poles along the outer radius of the hub to determine a new r2 for equation 2.1-16, see
Figure 2-9. This is accomplished simply by estimating the area of one half pole (as seen in the
FE mesh):
A p = h p (r3 − r2 ) , (2.1-21)

where hp is the half width of the pole and r3 and r2 are the pole-tip radius and the hub radius
respectively. Next the new value for “r2” can be found.
Atotal = Ahub + A poles = π rnew
2
(
− r12 ) (2.1-22)

Substituting equations for the various areas and solving for rnew we obtain.

(r )
8h p
rnew = r22 + − r2 (2.1-23)
π
3

This value is used in place of r3 when calculating the Ps due to the interference fit, equation 2.1-
16. This value for Ps is used in two ABAQUS models. The first is a model of the shaft with Ps
applied to the outer surface directed towards the center of the shaft (externally loaded cylinder).
The second is a model of the rotor with Ps applied to the inner surface (r1) directed radially
outward (internally loaded cylinder). These models are used to generate the stress field resulting
from the interference fit. This stress field is applied as an initial condition in the shaft-rotor
system FE (elastic and creep) models presented in this report, unless otherwise specified. Figure
2-10 shows the resulting stress field due to an interference fit of 0.0015’’ between a generic
titanium shaft and an iron-cobalt alloy rotor.

2.2 Finite Element Analysis


Once the ABAQUS input file (model.inp)2 has been generated using the program genmesh.F, the
FEA can be performed. This report contains an examination of both the elastic and creep

2
model.inp is the file name for the ABAQUS input file generated by the mesh generation program genmesh.F created by the author.

33
Figure 2-9 – Calculating r “new”

Figure 2-10 – Interference Fit Stress Profile

34
deformations. The elastic analysis was performed to establish a baseline with which to measure
the effects of the creep analysis. The creep analysis was done using the time hardening model
for an isotropic material. The ABAQUS time hardening model is:
ε! c = A(T )σ n t m 3 2.2-1

where A(T) is the stress coefficient which is a function of temperature, σ is the von Mises (or
effective) stress, t is time, and the n and m are exponents which are dependent on the material.
Fingers [3] found that the creep strain rate equation for the iron-cobalt alloy HA50HS is:
−Q
RT
A' e
ε! c = σ n' 2.2-2
T
Therefore
−Q

A(T ) =
A' e RT
T
n = n' 2.2-3
m=0

2.3 Post Processing of Results

ABAQUS generates the following result files; model.log, model.msg, model.dat, and model.fil4.
The log and msg files contain information pertaining to the convergence of the analysis. The fil
file is a result file designed to be portable between various programs and operating systems. All
of the stress and strain fringe plots presented in this report were generated using this file along
with PATRAN. The dat file is another result file that contains the user-requested output. The
post processor post.F5 uses this file to genenerate time histories of radial displacements (*.dv*
files), stress profiles (*_sts.exc file), and strain profiles (*_sts.exc file). For more details on
using post.F and the files it generates see Appendix A.

3
This equation represents the creep strain rate. Results output by ABAQUS are in the form of total strain which include the elastic strain similar
to equation 1.2-9, but neglecting the effects of transient effects and thermal expansion.
4
The ABAQUS results files have the extensions log, msg, dat, and fil. These extensions are added to the input filename. Therefore, if the input
filename is “model.inp” the results files will be model.log, model.msg, model.dat, and model.fil.
5
post.F is the post processing program created by the author.

35
3 Results and Discussion

The Air Force requirement for a material capable of maintaining structural integrity in high
temperature and high stress environments such as those found in jet engines is the motivation
behind this study. Discussions with Fingers [51] have led to the belief that the iron-cobalt rotor
would have to survive at 810K while spinning at 55,000 rpm for a lifetime of 5,000 hrs. For this
study the effects of start up and shut down have been ignored. Therefore, the life cycle model
assumes a steady state condition, a constant temperature profile; and a constant rotation speed for
the 5,000 hrs.

Dimensions of the shaft and the rotor were determined through discussions with Fingers [51].
The lamina is 0.006 inches thick. The shaft is comprised of a general titanium alloy 1 inch in
diameter with a wall thickness of 0.125 inches. The elastic modulus was taken as 16.8×106 psi.
The mass density was chosen to be 4.206×10-4 (lb s2/in4). The yield stress for titanium and its
alloys can range from 20 ksi to 120 ksi [46,47]. The rotor has a 1 inch diameter central hole, a
hub radius of 1.73 inches, and a pole-tip radius of 2.15 inches. Material properties for the iron-
cobalt alloy (HA50HS) rotor were determined in [3]. The elastic modulus is dependent on
temperature, as seen in Figure 3-1. The temperature dependence of the elastic modulus was
estimated using two linear fits. For temperatures between -65°F and 400°F the following
equation represents the elastic modulus:
E = (− 0.0012T + 25.729 )× 10 7 psi 3-1
where T is temperature measured in degrees Fairenheight. The equation for temperatures
between 600°F and 1200°F is:
E = (− 0.0104T + 29.727 )× 10 7 psi 3-2
The mass density and the yield stress were found to be 7.588×10-4 (lb s2/in4) and 180 ksi
respectively.

The first step in the analysis was to determine a mesh size that would adequately represent the
problem being analyzed. The first mesh examined contained 1560 elements, the second 2400
elements, and the third 3600 elements. There was a small difference between results computed

36
Figure 3-1 – Elastic Modulus vs. Temperature for HA50HS

37
with the 1560 and the 2400 element meshes but no noticeable difference in results computed
with the 2400 and the 3600 element meshes. Therefore, the 2400 element mesh was selected,
see Figure 3-2. The variables used to generate the mesh were:
NTHSEG = 60
NRADSEG = 25
NTBSEG = 10
NTRADPL = 10
The creep analysis requires input for the stress coefficient (A(T)), stress exponent (n), and time
exponent (m) in equation 2.2-1. A(T) is a function of temperature (T) as seen in equation 2.2-2
and Figure 3-3. The values of A(T) and n are also dependent on the creep type being analyzed.
Five creep models were defined in [3], see Table 1.2-3. Two of the creep types, A and AB1, are
examined in this report based on [51].

Two failure criteria were decided upon based on [51]. The first is the pole-tip-rubbing criterion.
This failure occurs if the rotor pole-tip deformations are large enough for it to come into contact
with the stator poles. This failure will occur at pole-tip deformations between 0.02 and 0.04
inches. The second failure criterion involves the interference fit. This failure will occur when
the radial stress (σr) at the outer radius of the shaft is relaxed to zero in the shaft, indicating a
separation of the rotor from the shaft.

3.1 Results from the Elastic Analysis

An elastic analysis was performed to establish a baseline with which to compare the creep
results. This analysis included a temperature range from 727K to 810K (nearly 850°F to
1000°F) and rotation speed (ω) ranging from 45,000 to 65,000 rpm. The effect of the
interference fit is also studied by varying the interference fit (d) from 0.0015 to 0.003 inches.
The temperature is assumed to be constant in each lamina.

38
Figure 3-2 – 2400 Element Mesh

Figure 3-3 – Stress Coefficient vs. Temperature for HA50HS

39
3.1.1 Single Lamina Results

Based on the results of the elastic analysis one can conclude that temperatures in the range of
interest for this study do not significantly effect the elastic deformations and stresses. As
depicted in Figure 3-4, a rotor with an interference fit equal to 0.0015 inches subjected to a
rotation speed of 45,000 rpm shows almost the same stress field for both 727K and 810K. The
radial displacement profile, Figure 3-5, indicates the same phenomenon. The only temperature
dependent variable for the elastic cases is the elastic modulus, which exhibits only a 7.5%
change in magnitude over this temperature range. Therefore, it is reasonable to find that
temperature does not significantly effect the elastic deformations.

One of the advantages of the FEA can be seen in Figures 3-4 and 3-5. Past analyses assumed
that pole loading could be smoothed over the outer hub radius, and thus would yield results
which were independent of the circumferencial position [30]. The reality is that the poles would
cause concentrated loading on the outer radius of the hub resulting in stress and deformation
fields that exhibit a dependence on the angular position. Figure 3-5 shows this dependence. The
effects of the poles are clearly seen near the outer radius of the hub where the data points for a
given radial position show a significant disparity6 for the radial displacement. This disparity is
not as apparent near the inner radius of the hub where the data points for a given radial position
nearly fall on top of each other. Figures 3-6 and 3-7 clearly depict the dependence on the
tangential position. The gradient lines for the strains exhibit an almost circular pattern near the
shaft indicating an independence from the tangential position. The gradient lines near the hub or
rotor poles do not follow a circular pattern showing a dependence on the tangential position
resulting from the rotor pole loading. It can be concluded that the poles induce a dependence on
the tangential position throughout the shaft-rotor system and that magnitude is dependent on the
proximity to the poles.

6
Since there are many elements at each radial position, and the poles cause a dependence on the circumferential position the stress and radial
displacement profiles have multiple data points at each radial position causing the disparity in values

40
Figure 3-4 – Elastic Stress Profiles for 727K and 810K for d=0.0015’’, ω=45,000rpm

Figure 3-5 – Elastic Radial displacement Profiles for 727K and 810K for d=0.0015’’, ω=45,000rpm

41
Figure 3-6 – Elastic Von Mises Strain Fringe Plot for d=0.0015’’, T=727K, ω=45,000rpm

Figure 3-7 – Elastic Von Mises Strain Fringe Plot for d=0.0015’’, T=810K, ω=45,000rpm

42
The results indicate that the rotation speed (ω) greatly affects the elastic stress and radial
displacement profiles. Figures 3-8 shows the effects of ω on a shaft-rotor system at 727K with d
equal to 0.0015 inches. The pole-tip radial displacements for both cases do not exceed the
failure criteria defined in section 3. The pole-tip radial displacement for the 45,000 rpm case is
approximately 1.6 × 10 −3 inches and 2.3 × 10 −3 inches for the 55,000 rpm case. It is clear that for
the 45,000 rpm case the interference fit remains intact. This is not the case for the 55,000 rpm
case where the radial stresses in the shaft have become tensile. This indicates that the elastic
stresses at 727K and 55,000 rpm are severe enough to cause the rotor to become separated from
the shaft. It is clear, without performing the creep analysis, that the desired operating
environment of 810K, 55,000 rpm, and a life cycle of 5,000 hrs is not possible for a constant
temperature rotor with an interference fit equal to 0.0015 inches, since the elastic case
experiences failure at 727K at 55,000 rpm. The rest of this report focuses on determining the
safe operational environment for the given material and configuration.

Temperatures of 760K, 780K, and 795K were also examined. The stress profiles for the 45,000
rpm is presented in Figures 3-9 through 3-11. In all three cases the stress profile is nearly
identical. The radial stress is zero at the shaft hole and decreases to about –2,000 psi at a radius
of 0.5 inches or the rotor hole. This trend depicts the effect of the interference fit. The radial
stress then increases to a maximum value of about 18,000 psi at 1 inch in radius. The radial
stress then decreases slightly until it reaches the hub radius of 1.73 inches. At this radius the
average radial stress value is nearly 10,000 psi. There is a split in the radial stress, which occurs
near the hub radius. This indicates the pull of the rotor poles. The tangential stress begins at a
value near –10,000 psi at the shaft hole and then increases slightly to about –8000 psi at the rotor
hole. At this radius there is a discontinuity. The tangential stress jumps to a maximum value of
about 54,000 psi. From here it decreases almost asymptotically to an average value of about
20,000 psi at the outer surface of the rotor. Again the effects of the poles can clearly be seen in
the disparity of the radial displacement and stress values at a given radial position. Unlike
previous studies [30] where the rotor poles were neglected as part of the geometry, this study
includes the poles to obtain more realistic stress and radial displacement profiles. The result is
that the rotor poles induce a dependence on the tangential position, which is reflected in both the

43
Figure 3-8 – Elastic Stress Profiles for 45,000 rpm and 55,000rpm for d=0.0015’’, T=727K

Figure 3-9 – Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=45,000rpm, T=760K

44
Figure 3-10 – Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=45,000rpm, T=780K

Figure 3-11 – Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=45,000rpm, T=795K

45
disparities of the radial displacements and the stresses for a given radial position near the hub
radius. There is no interference fit failure for these models. Figure 3-12 shows that there is no
pole-tip rubbing as a result of elastic deformations at these temperatures.

The rotation speed of 55,000 rpm was also examined at a temperature range of 727K to 810K, as
depicted in Figures 3-13 through 3-16. As in the 727K case, the rotation speed of 55,000 rpm
results in interference fit failure. Here again the stress profiles for the four temperature cases are
identical. The radial stress begins at zero at the shaft hole radius. Unlike the previous lower rpm
cases, these cases reveal a positive increase in the radial stress from the shaft to rotor hole. This
positive radial stress indicates that the interference fit no longer holds the rotor to the shaft. If
the rotor was not “tied” to the shaft then the rotor and the shaft would have become separated
and the system would experience interference fit failure.

The results of the elastic analysis thus far suggest that the interference fit is the critical design
parameter. Therefore, it was decided to examine the effect of increasing the interference fit (d).
The interference fit (d) was doubled from 0.0015 to 0.003 inches, thereby doubling the
equivalent pressure (Ps) from equation 2.1-16. The result of this increase in the interference fit is
that the rotor will stay attached to the shaft at higher rotation speeds for the desired lifetime of
5,000 hrs. Previously it was shown that at d equal to 0.0015 inches, a temperature of 727K, and
a rotation speed (ω) of 55,000 rpm, the interference fit failure would occur, see Figure 3-8. If d
is increased to 0.003 inches this failure does not occur under the same conditions, see Figure 3-
17. This result can also be seen at temperatures of 760K, 780K and 810K, see Figures 3-18
through 3-20. The increase in the interference fit has resulted in a significant change in the
elastic stress profile though there is still an apparent independence from temperature. The radial
and tangential stress profiles follow the same trend as seen in the d equal to 0.0015 inch and ω
equal to 45,000 rpm cases seen in Figures 3-9 through 3-11, though the maximum values have
changed as a result of the new interference fit value. The maximum negative value for the radial
stress is now about –10,000 psi indicating a stronger interference fit than in the lower rpm and
interference fit cases. The radial stress then increases to a maximum value of about 23,000 psi
again at about 1 inch in radius. This represents an increase of about

46
Figure 3-12 – Elastic Pole-tip Radial displacement vs. Temperature for ω=45,000rpm

Figure 3-13 - Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=55,000rpm, T=760K

47
Figure 3-14 - Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=55,000rpm, T=780K

Figure 3-15 - Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=55,000rpm, T=795K

48
Figure 3-16 - Elastic Stress Profile for Shaft-Hub System d=0.0015’’, ω=55,000rpm, T=810K

Figure 3-17 - Elastic Stress Profile for Shaft-Hub System d=0.003’’, ω=55,000rpm, T=727K

49
Figure 3-18 - Elastic Stress Profile for Shaft-Hub System d=0.003’’, ω=55,000rpm, T=760K

Figure 3-19 - Elastic Stress Profile for Shaft-Hub System d=0.003’’, ω=55,000rpm, T=780K

50
Figure 3-20 - Elastic Stress Profile for Shaft-Hub System d=0.003’’, ω=55,000rpm, T=810K

51
5,000 psi from the previous models. The tangential stress is about –40,000 psi at the shaft hole,
almost 4 times greater than the d equal to 0.0015 inch and ω equal to 45,000 rpm models. The
maximum value for the tangential stress has increased from about 30,000 psi to about 84,000 psi.
This value decreases to about 30,000 psi at the hub radius which is about 10,000 psi greater than
that in the previous models.

There is also no failure due to pole-tip-rubbing at these temperatures as indicated in Figure 3-21.
Comparing the 45,000 rpm cases (d=0.0015’’ and d=0.003’’) it is found that the size of the
interference fit has no significant effect on the elastic radial displacements of the poles, see
Figure 3-12. It is again apparent that the rotation speed (ω) has a more significant effect on the
elastic radial displacements, see Figure 3-22. The pole-tip radial displacement is about
1.55 × 10 −3 inches at 45,000 rpm and increases to about 2.80 × 10 −3 inches at 60,000 rpm. These
values are one tenth of the magnitude necessary to initiate pole-tip rubbing.

52
Figure 3-21 – Elastic Pole-tip Radial displacement vs. Temperature for d=0.003’’, ω=55,000rpm

Figure 3-22 – Elastic Pole-tip Radial displacement vs. RPM for d=0.003’’, T=727K

53
3.2 Results from the Creep Analysis

The creep analysis was performed in three phases. The first phase was the study of constant
temperature rotor laminas tied to a shaft of equal thickness7. A temperature range of 727K to
810K was examined. Since the material creep properties were unknown for the shaft (because
the shaft material is unknown) the assumption was made that these properties were equal to those
of the HA50HS rotor material. A range in rotation speed (ω) was also examined for each value
of temperature in order to determine the maximum allowable rotation speed (ω). Finally, the
effect of varying the interference fit (d) from 0.0015 to 0.003 inches was examined. All of this
analysis was performed for both the creep type A and AB1 models.

The second phase of the creep analysis examined a single rotor lamina subjected to a linear radial
variation in temperature tied to a shaft of equal thickness. Three temperature ranges were
examined. The first assumed that the inner radius of the rotor (rotor hole) was at 588K (about
600°F) and the pole-tips were at 810K (about 1000°F). The second model assumed a hotter
environment where the rotor hole was at 727K (about 850°F) and the pole-tips were again at
810K. Both of these temperature models can be thought of as realistically representing the type
of environment the rotor would be subjected to. The third model was the inverse of model one
in that the rotor hole temperature was assumed to be 810K and the pole-tip temperature was
assumed to be 588K. This model was examined out of curiosity and the results were deemed
interesting enough to present in this report. This phase only examined the creep type A material
and d equal to 0.0015’’.

The final phase of the creep analysis examined two rotor lamina with an axial variation in
temperature. Each lamina is modeled at a constant but different temperature. The stack is held
together by an axial compressive force. For these models the rotor lamina is 20 times thicker
(0.12 inches) than that in phase one and two. The increase in thickness results in a larger
temperature difference between the two laminae than would occur for the thinner (0.006 inches)

7
The shaft was modeled as one lamina thick to prevent any distortion of results due to the edge effect.

54
models. The surface interaction between the laminae is modeled as Coulomb (static) friction.
These models provide a first step towards the understanding of the behavior of the entire rotor
stack..

3.2.1 Single Lamina Results – Constant Temperature Models

3.2.1.1 727K

The 727K constant temperature model was examined for the case where d is 0.0015 inches, the
creep type is A for a range of rotation speed (ω) from 35,000 to 50,000 rpm in 5,000 rpm
increments. The results indicate that the shaft-rotor system does not experience failure prior to
the 5,000 hr life cycle between 35,000 and 45,000 rpm see Figures B-1 to B-38. The stress
profiles for these three cases show little change over time indicating that they are not
experiencing a large amount of creep. The profiles follow the same trend as seen in the elastic
results, as expected, since there is little creep to affect the results. The 35,000 rpm case, Figure
B-1, shows that the radial stress reaches a maximum negative value of –5,000 psi in the shaft,
then increases to 10,000 psi at about 1 inch in radius. The tangential stress begins at about –
29,000 psi at the shaft hole at 0.257 seconds, which then increases to about –26,000 psi at 5,000
hrs. The maximum tangential stress occurs at the rotor hole both at 0.257 seconds and at 5,000
hrs. The value is about 39,000 psi at 0.257 seconds, which decreases to about 34,000 psi at
5,000 hrs. Both the radial and the tangential stress profiles appear unchanged with time at radial
positions greater than about 0.6 inches. The lack of creep deformations can be seen in the radial
displacement vs. time plot in Figure B-1. There is only a minute increase in the radial
displacement over the time range. This increase represents the creep deformations. Figure B-2
shows the effect of increasing the rotation speed by 5,000 rpm to 40,000 rpm. The stress profiles
have the same trend only the numeric values have changed. The radial stress only falls to about
–5,000 psi in the shaft indicating that a greater amount of the interference fit stress has been
relaxed, bringing the shaft-rotor system closer to failure. The radial stress attains a maximum
value of about 12,000 psi at 1 inch in radius. The maximum tangential stress has increased to

8
Due to large number of cases run the majority of the results are presented in Appendices B and C.

55
about 45,000 psi, which is about 6,000 psi greater than that for the 35,000 rpm case. The radial
displacement profile shows that the shaft hole is displacing a greater amount than the rotor hole,
indicating that the system is becoming more compressed at the interior radii. The 45,000 rpm
case again shows the same trends as the previous two cases, except it is apparent that for this
rotation speed the system comes closer to failure, indicated by the near zero values for the radial
stress at the outer surface of the shaft. This second increase in rotation speed increased the radial
stress values in the shaft as well as increased the maximum tangential stress value to nearly
55,000 psi at the rotor hole.

The results for 50,000 rpm indicate interference fit failure (a separation of the rotor from the
shaft) between 75 hrs and 5,000 hrs, Figure 3-23. The large range in probable time of failure is
due to the frequency of output during the FEA. The radial displacement profile can be seen in
Figure 3-24. This profile resembles the profile of a spinning disk free to deform at its inner
radius. The obvious difference is the disparity in the radial displacements at radial positions near
the outer hub of the rotor. This disparity is the tangential position dependence resulting from the
loads induced by the rotor poles. This “pole effect” is more clearly depicted in Figure 3-25.
This plot shows the distribution of the Von Mises strain over the entire shaft-rotor system. The
gradient pattern in and near the shaft is nearly circular, indicating an independence from the
tangential position. The gradient pattern becomes less circular as the radial position approaches
the outer hub, indicating a dependence on the tangential position. This pattern shows the poles
acting as concentrated loads acting on the outer portion of the FE model. These loads cause the
part of the rotor near the poles to experience greater deformations than the rest of the rotor.

The time history of the radial displacements is depicted in Figure 3-26. Time equal to zero
indicates elastic radial displacements. It is apparent that the shaft-rotor system does not undergo
a significant amount of creep radial displacements prior to failure.

The effect of the rotation speed (ω) on the pole-tip radial displacements can be seen in Figure 3-
27. The bottom line indicates the radial displacement of the pole-tips after 0.2 seconds, or
essentially the elastic radial displacements. The top line represents the radial displacements after
5,000 hrs. The symbol IFF found on the 5,000 hr line between 45,000 rpm and 50,000 rpm

56
Figure 3-23 – Shaft-Hub Radial Stress profile for d=0.0015’’, ω=50,000rpm, T=727K, Type A, ATi=AFeCo9

Figure 3-24 – Shaft-Hub Radial displacement Profile for d=0.0015’’, ω=50,000rpm, T=727K, t=75hrs, Type
A, ATi=AFeCo

9
ATi stands for the stress coefficient for the titanium shaft. AFeCo stands for the stress coefficient for the iron-cobalt alloy rotor.

57
ω=50,000rpm, T=727K, t=75hrs, Type A,
Figure 3-25 – Von Mises Strain Fringe Plot for d=0.0015’’,ω
ATi=AFeCo

Figure 3-26 – Radial displacement vs. Time for d=0.0015’’, ω=50,000rpm, T=727K, Type A, ATi=AFeCo

58
Figure 3-27 – Pole-tip Radial displacement vs. RPM for d=0.0015’’, T=727K, Type A, ATi=AFeCo

59
indicates that the rotor experiences interference fit failure prior to 5,000 hrs at a rotation speed
(ω) between 45,000 and 50,000 rpm. This symbol also means that the 5,000 hr value for radial
displacement is not valid at 50,000 rpm (even though it is plotted) since this value indicates
5,000 hrs of analysis but it does not take into account failure10. It is clear that at 727K for the
prescribed material variables and conditions the shaft-rotor system is not experiencing a
significant amount of creep deformations. The elastic deformations are dominant in this regime.
Furthermore, the pole-tip deformations are on the order of one tenth the magnitude necessary to
cause pole-tip rubbing. It appears that the critical design criterion is the interference fit; this
conclusion was also reached after examining the elastic results.

The analysis described above was performed for the case where the interference fit (d) is 0.003
inches. As indicated in the elastic analysis, increasing the interference fit (d) effectively
increases the rotation speed range at which the shaft-rotor system can survive for 5,000 hrs. The
interference fit failure now occurs between 55,000 rpm and 60,000 rpm, as depicted in Figure 3-
28. The interference fit failure occurs at 60,000 rpm at some time between 2292 hrs and 5,000
hrs, Figure 3-29. It is clear from Figure 3-28 that the elastic deformations are not affected by the
increased interference fit (as observed in the elastic analysis). Increasing the interference fit (d)
does allow for an increase in rotation speed (ω) resulting in an increase in the creep
deformations, though the elastic deformations still comprise a significant amount of the overall
deformations. The radial displacement profile shows the same trend as the d equal to 0.0015
inch cases, see Figure 3-30., and as expected the increased rotation speed results in an increase in
radial displacement magnitudes. The tangential position dependence resulting from the “pole
effect” is again apparent. The “pole effect” is made clearer in the strain fringe plot in Figure 3-
31, where the gradient lines in the hub deviate from the nearly circular patterns found in the
shaft.

10
The FEA does not stop when interference fit failure occurs. The shaft is “tied” to the rotor; therefore, it will not actually separate during the
FEA. Interference fit failure is only determined after examining the results and determining where the radial stress in the shaft has exceeded zero.

60
Figure 3-28 – Pole-tip Radial displacement vs. RPM for T=727K, Type A, d=0.0015’’ and d=0.003,
ATi=AFeCo’’

Figure 3-29 – Shaft-Hub Radial Stress Profile for d=0.003’’, T=727K, ω=60,000rpm, Type A, ATi=AFeCo

61
Figure 3-30 – Shaft-Hub Radial displacement Profile for d=0.003’’, ω=60,000rpm, T=727K, t=2292hrs, Type
A, ATi=AFeCo

ω=60,000rpm, T=727K, t=2292hrs, Type A,


Figure 3-31 – Von Mises Strain Fringe Plot for d=0.003’’,ω
ATi=AfeCo

62
Figures B-4 through B-6 contain the stress and deformation profiles for the 45,000 rpm, 50,000
rpm, and 55,000 rpm cases also run at this temperature. The 45,000 rpm case shown in Figure
B-4, when compared to the case in Figure B-3, clearly shows the effect of increasing the
interference fit. The radial stress in the shaft is about –11,000 psi after 5,000 hrs, about ten times
greater than the d equal 0.0015 inch case. The increase in d results in a greater variance in the
tangential stress profile between 0.257 seconds and 5,000 hrs, Figure B-3. The maximum value
drops from about 55,000 psi to about 45,000 psi, a percent change of 20%. In Figure B-4 this
value drops from 63,000 psi to 43,000 psi or a drop of about 30%. This can be attributed to the
increase in the stress values as a result of the interference fit increase. The creep rate is highly
dependent on the stress as seen in equation 2.2-1. Increasing the rotation speed increases the
radial stress at the shaft hole from about –6,500 psi to about –5,500 psi for the 50,000 and 55,000
rpm cases respectively at 0.257 seconds, see Figures B-5 and B-6. The maximum tangential
stress at 5,000 hrs increases from 79,000 to about 84,000 psi for the same increase in rotation
speed. The 5,000 hr pole-tip radial displacements increase from 1.69 × 10 −3 to 2.28 × 10 −3 and
finally to 2.75 × 10 −3 inches for the 45,000, 50,000, and 55,000 rpm models, see Figures B-4
through B-6.

The time history of radial displacements is depicted in Figure 3-32. This plot indicates a greater
magnitude of creep radial displacements prior to failure than in the d equal 0.0015 inch model
depicted in Figure 3-26. This is most likely caused by the increase in survivable rotation speed
that is a result of the increase in the interference fit. Comparison of the d equal to 0.0015 and
0.003 inch cases at ω equal to 45,000 rpm in Figure 3-28 indicates that for similar test cases
there is only a small increase in creep deformations resulting from an increase in d.

All of the analyses described in this section were also performed using material variables from
the creep type AB1 model outlined in [3]. The results indicate that there is no appreciable
difference in either the stress or deformation fields at the temperature of 727K. This result is to
be expected since the creep type A and AB1 models are very similar. This is clearly seen in
Figure 3-2 where the stress coefficient (A(T)) for the two models is essentially the same until
temperatures of 780K. Therefore, it is reasonable to assume that any significant difference

63
Figure 3-32 – Radial displacement vs. Time for d=0.003’’, T=727K, ω=60,000rpms, Type A, ATi=AFeCo

Figure 3-33 – Pole-tip Radial displacement vs. RPM for T=727K, Type AB1, d=0.0015’’ and d=0.003’’,
ATi=AfeCo

64
between the two models would manifest itself at temperatures at or above 780K. The pole-tip
radial displacements at various rotation speeds (ω) for the interference fit (d) equal to 0.0015 and
0.003 inches models can be seen in Figure 3-33. The stress and radial displacement profiles for
d equal to 0.0015 inches at various ω can be found in Figures C-1 through C-4 and the profiles
for d equal to 0.003 inches are depicted in Figures C-5 through C-8.

3.2.1.2 760K

The 760K constant temperature model was examined for the case where d is 0.0015 inches, the
creep type is A for a range of rotation speeds (ω) from 30,000 rpm to 40,000 rpm in 5,000 rpm
increments. The shaft-rotor system does not fail prior to the 5,000 hr time limit at rotation
speeds (ω) between 30,000 and 35,000 rpm, see Figures B-7 and B-8. The increase in
temperature has resulted in an increase in creep related deformations. This is apparent even at
the low rotation speed of 30,000 rpm in Figure B-7. Even though this figure contains results for
a rotation speed 5,000 rpm less than seen in the 727K models, the stress profiles show a definite
change with time not previously seen. The radial stresses at the rotor hole increase from about –
9000 to –5,000 psi, while the maximum tangential stress decreases from about 31,000 to 19,000
psi between 0.257 seconds and 5,000 hrs. The radial stress at radial positions greater than 0.6
inches do not show any change over time as was seen in the previous temperature profile. The
tangential stress profile shows no change at radial positions greater than about 0.8 inches, this
represents a shift of about 0.2 inches from the 727K temperature model. The higher temperature
profile results in greater creep deformation as can be seen in the radial displacement vs. time plot
in Figure B-7. Figure B-8 contains data resulting from an increase of 5,000 rpm to 35,000 rpm.
The increase in rotation speed has brought the system closer to failure as indicated by the nearly
zero radial stress in the shaft at 5,000 hrs in Figure B-8. The radial stress increased from about –
6,000 to –2000 psi between 0.257 seconds and 5,000 hrs. There is also a significant change in
the tangential stress in the shaft. At 0.257 seconds the tangential stress in the shaft begins at
about –28,000 psi at the shaft hole and increases to about –21,000 psi at the rotor hole. At 5,000
hrs the tangential stress has increased to an almost constant –9,000 psi through out the shaft. The
maximum tangential stress at the rotor hole decreases from about 38,000 to about 22,000 psi

65
between 0.257 seconds and 5,000 hrs. This represents about a 42% drop, which is close to the
39% drop seen in the 30,000 rpm case.

Failure occurs some time between 2,052 and 5,000 hrs at 40,000 rpm as indicated in Figure 3-34.
The radial stress increases from zero to about –5,000 psi in the shaft for the 0.257-second data.
This trend clearly shows that the interference fit is intact. The 5,000 hr case shows that the radial
stress in the shaft has relaxed to a positive value indicating tension or a separation of the rotor
from the shaft. The radial stress in the rotor increases from the inner radius (rotor hole) to a
maximum value of about 13,000 psi (for the 0.257 second data) at around 1 inch in radius. The
radial stress then decreases as one approaches the outer radius of the hub. This trend is similar to
that found from the spinning disk case with a free (unclamped) inner radius as discussed in the
previous section. The radial displacement profile and strain fringe plot are in Figures 3-35 and
3-36 respectively. The time history of radial displacements is depicted in Figure 3-37. Again,
the effects of the rotor poles are clear as is the fact that the creep radial displacements do not
greatly exceed the elastic (time equal to zero) radial displacements prior to failure.

The increase of temperature from 727K to 760K has reduced the survivable rotation speed range
from 45,000 to 50,000 rpm to 35,000 to 40,000 rpm. Failure occurs as a result of interference fit
failure in both cases. Figure 3-38 shows once again that the pole-tip radial displacements are not
approaching the magnitude necessary to cause pole-tip rubbing.

The effect of increasing the interference fit (d) to 0.003 inches was examined. The rotation
speed range examined was from 35,000 rpm to 45,000 rpm in 5,000 rpm increments. Both the
35,000 and 40,000 rpm case survived for the required 5,000 hrs, see Figures B-9 and B-10. The
increase in the interference fit results in a decrease in the radial stress at the rotor hole from –
6,000 to about –18,000 psi between the d equal to 0.0015 and 0.003 inch 35,000 rpm cases, see
Figure B-8 and B-9. In Figure B-9 the radial stress increases to about –5,000 psi at 5,000 hrs at
the rotor hole. The tangential stress shows a large change between the 0.257 seconds and 5000hr
profiles. At 0.257 seconds the tangential stress increases from –81,000 to –64,000 psi in the
shaft. The tangential stress decreases from –18,000 to –19,000 psi in the shaft at 5,000 hrs. The

66
Figure 3-34- Shaft-Hub Stress Profile for d=0.0015’’, T=760K, ω=40,000rpm, Type A, ATi=AFeCo

67
Figure 3-35 – Shaft-Hub Radial displacement Profile for d=0.0015’’, ω=40,000rpm, T=760K, t=2052hrs, Type
A, ATi=AFeCo

ω=40,000rpm, T=760K, t=2052hrs, Type A,


Figure 3-36 – Von Mises Strain Fringe Plot for d=0.0015’’,ω
ATi=AFeCo

68
Figure 3-37 – Radial displacement vs. Time for d=0.0015’’, T=760K, ω=40,000rpms, Type A, ATi=AfeCo

Figure 3-38 – Pole-tip Radial displacement vs. RPM for T=727K and T=760K, Type A, d=0.0015’’,
ATi=AfeCo

69
maximum tangential stress at the rotor hole decreases from about 50,000 to about 21,000 psi
between 0.257 seconds and 5,000 hrs. The radial position at which the radial stress no longer
shows any difference with respect to time is now about 1 inches. The radial position at which
this occurs for the tangential stress is about 1.18 inches. This can be attributed to the increase in
overall creep due to the increase in temperature and rotation speed. Increasing the rotation speed
to 40,000 rpm nearly causes the system to experience failure as indicated by the nearly zero
radial stresses at 5,000 hrs in the shaft in Figure B-10. The radial stress increases from about –
18,000 to about –1000 psi in the shaft between 0.257 seconds and 5,000 hrs for the 40,000 rpm
case. The tangential stress also approaches zero in the shaft increasing from about –76,000 to
about –3,000 psi at the shaft hole between 0.257 seconds and 5,000 hrs. The maximum
tangential stress at the rotor hole decreases from 59,000 to 25,000 psi between 0.257 seconds and
5,000 hrs. The pole-tip deflections increase from about 1.25 × 10 −3 to about 1.95 × 10 −3 inches
over the given time period as seen in the radial displacement vs. time plot in Figure B-10.

Interference fit failure occurs very near 2,958 hrs at the rotation speed (ω) of 45,000 rpm, see
Figure 3-39. The range of failure has increased by 5,000 rpm due to the increase in the
interference fit (d), see Figure 3-40. The same increase was found in the 727K constant
temperature case. The radial displacement profile, strain fringe plot, and time history of radial
displacements are found in Figures 3-41 through 3-43 respectively.

All of the analyses described in this section were also performed using material variables from
the creep type AB1 model outlined in [3]. The pole-tip radial displacements at various rotation
speeds (ω) for the interference fit (d) equal to 0.0015 and 0.003 inches models can be seen in
Figure 3-44. The stress and radial displacement profiles for the d equal to 0.0015 inches at
various ω can be found in Figures C-9 through C-11 and the profiles for d equal to 0.003 inches
are depicted in Figures C-12 through C-14.

70
Figure 3-39 - Shaft-Hub Stress Profile for d=0.003’’, T=760K, ω=45,000rpm, Type A, ATi=AFeCo

Figure 3-40 – Pole-tip Radial displacement vs. RPM for T=760K, Type A, d=0.0015 and d=0.003’’’’,
ATi=AfeCo

71
Figure 3-41 – Shaft-Hub Radial displacement Profile for d=0.003’’, ω=45,000rpm, T=760K, t=2958hrs, Type
A, ATi=AfeCo

ω=45,000rpm, T=760K, t=2958hrs, Type A,


Figure 3-42 – Von Mises Strain Fringe Plot for d=0.003’’,ω
ATi=AFeCo

72
Figure 3-43 – Radial displacement vs. Time for d=0.003’’, T=760K, ω=45,000rpms, Type A, ATi=AFeCo

Figure 3-44 – Pole-tip Radial displacement vs. RPM for T=760K, Type A, d=0.0015 and d=0.003’’,
ATi=AFeCo

73
3.2.1.3 780K

The 780K constant temperature model was examined for the case where d is 0.0015 inches, the
creep type is A for a range of rotation speeds (ω) from 25,000 to 35,000 rpm. The results
indicate that the shaft-rotor system does not experience failure prior to the 5,000 hr limit between
25,000 and 30,000 rpm, see Figures B-11 to B-13. The time history plot in Figure B-11 shows
an interesting trend for the initial radial displacements. There is a slight decrease in the amount
of radial displacement for a small period of time before the trend shows the typical increase of
radial displacement with time. This trend will be discussed in greater detail later in this report.
The radial stress at the rotor hole increases from about –9,000 to about –3,000 psi for the 25,000
rpm case between 0.257 seconds and 5000 hrs, as depicted in Figure B-11. The tangential stress
at the shaft hole increases from –40,000 to –10,000 psi. The maximum tangential stress at the
rotor hole decreases from about 25,000 to about 11,000 psi between 0.257 seconds and 5000 hrs.
The radial displacement profile indicates an almost uniform radial displacement over the shaft
followed by only a slight dip in the interior of the rotor. This trend indicates that in this
environment the shaft and interior of the rotor do not undergo the same compression as seen in
previous models. This is clearly depicted in the radial displacement vs. time plot where the rotor
hole radial displacements are slightly larger than the shaft hole radial displacements. Figure B-
12 contains data from the 27,000 rpm model. The stress profiles indicate only a small change
from the 25,000 rpm model. The increase in rotation speed resulted in an increase in pole-tip
radial displacement at 5,000 hrs from about 6.1× 10 −4 to 7.9 × 10 −4 inches between 25,000 and
27,000 rpm. The increase in speed also resulted in a greater degree of compression in the shaft
and interior of the rotor as indicated by the dip in radial displacements from the shaft hole to 0.7
inches. It is clear, however, from the radial displacement vs. time plot that the compression is
not occurring in the shaft. Figure B-13 contains data from the 30,000 rpm case. Comparing
results from Figures B-13 and B-7 it is clear that the increase in temperature from 760 to 780K
results in a greater effect of creep on the overall deformations. The radial stress at the rotor hole
increases from about –3,000 to –10,000 psi between 760 and 780K for 30,000 rpm. The rotor
hole tangential stress decreases from 19,000 to about 15,000 psi at 5,000 hrs for the same
increase in temperature. The pole-tip radial displacement at 5,000 hrs increases from 8.0 × 10 −4
to 1.1× 10 −3 inches for the increase in temperature from 760 to 780K. The radial displacement

74
vs. time plot in Figure B-13 indicates that the shaft is experiencing compression as indicated by
the shaft hole displacing a greater magnitude than the rotor hole

The 35,000 rpm results indicate interference fit failure between 943 and 5,000 hrs, see Figure 3-
45. The radial displacement profile, strain fringe plot, and time history of radial displacements
for this case are depicted in Figures 3-46 through 3-48 respectively. The increase in temperature
from 760K to 780K has reduced the survivable rotation speed range from 35,000 to 45,000 rpm
to 30,000 to 35,000 rpm. For both temperatures the failure was the result of the rotor becoming
separated from the shaft (interference fit failure). The pole-tip radial displacements experienced
prior to failure were an order of magnitude smaller than that required to induce pole-tip rubbing,
see Figure 3-49.

The effect of increasing the interference fit (d) to 0.003 inches was examined. The rotation
speed range examined was from 25,000 to 35,000 rpm in 5,000 rpm increments. Both the 25,000
and 30,000 rpm case survived the required 5,000 hrs, see Figures B-14 and B-15. Interference fit
failure occurs between 642 and 4,672 hrs at a rotation speed (ω) of 35,000 rpm, see Figure 3-50.
Unlike the previous temperatures examined, the range of survivable rotation speed has not
changed as a result of the increase in d. A closer examination of the results from the 727 and
760K constant temperature models reveals an interesting trend. The 727K case showed an
increase in survivability from 45,000 to 50,000 rpm when the interference fit (d) was increased.
This case also showed that at d equal to 0.0015 inches and ω equal to 50,000 rpm, failure occurs
between 75 and 5,000 hrs. The case of d equal to 0.003 inches and ω equal to 55,000 rpm
indicated failure between 2,292 and 5,000 hrs. This data leads to the conclusion that for the d
equal 0.0015 inches the maximum survivable rotation speed (ω) is closer to 50,000 rpm than it is
to 55,000 rpm. On the other hand, for the d equal to 0.003 inches case the maximum is closer to
60,000 rpm than it is to 55,000 rpm. An examination of the 760K constant temperature case
indicates that the amount of rotation speed (ω) gained from increasing the interference fit is
smaller than it was for the 727K case. For the d equal to 0.0015 inches case, failure occurs
between 2,052 and 5,000 hrs at 40,000 rpm. Failure then occurs between 2,958 and 5,000 hrs at
45,000 rpm for the d equal to 0.003 inches case. The 727K results indicate a gain in maximum
survivable rotation speed of around 7,000 to 10,000 rpm from the increase in d. The 760K

75
Figure 3-45 - Shaft-Hub Stress Profile for d=0.0015’’, T=780K, ω=35,000rpm, Type A, ATi=AfeCo

Figure 3-46 – Shaft-Hub Radial displacement Profile for d=0.0015’’, ω=35,000rpm, T=780K, t=943hrs, Type
A, ATi=AFeCo

76
ω=35,000rpm, T=780K, t=943hrs, Type A,
Figure 3-47 – Von Mises Strain Fringe Plot for d=0.0015’’,ω
ATi=AFeCo

Figure 3-48 – Radial displacement vs. Time for d=0.0015’’, T=780K, ω=35,000rpms, Type A, ATi=AfeCo

77
Figure 3-49 – Pole-tip Radial displacement vs. RPM for T=780K, 760K, and 727K, Type A, d=0.0015’’,
ATi=AfeCo

Figure 3-50 - Shaft-Hub Stress Profile for d=0.003’’, T=780K, ω=35,000rpm, Type A, ATi=AfeCo

78
results indicate a gain in around 5,000 to 7,000 rpm from the same increase in d. This trend
indicates that as the temperature is increased the benefit from increasing the interference fit (d) is
less. This is further supported by the data from the 780K case where there appears to be little or
no significant increase in survivable rotation speed as a result of increasing d, see Figure 3-51.

The 780K and d equal to 0.003 inches data reveal another fascinating phenomenon at the lower
rotation speeds, see Figure B-14. The radial displacement vs. time plot does not reveal the
previous trend of increasing radial displacement with time. After an initial positive elastic radial
displacement the trend decreases with time until about 1,670 hrs where the radial displacements
begin to increase with time. This indicates that for a period of time the shaft-rotor system is
experiencing negative radial displacements. This actually causes the rotor pole to initially pull
away from the stator poles. The final deformation for the rotor poles turns out to be slightly less
than the initial elastic deformation.

The data can be interpreted as follows: The shaft-rotor system undergoes elastic deformation due
to the loads caused by the rotation at time zero. After time zero, the creep model dictates the
deformation model. The model used is the power law model where strain is dependent on stress.
For the 780K, d equal to 0.003 inches, and ω equal to 25,000 rpm, case the initial stress field is
dominated by the stresses due to the interference fit. In other words, the stress resulting from the
rotation loads does not quickly relax or overcome the interference fit stresses as previously seen.
Therefore, the stress field initially dictates a negative radial displacement. At a later time the
stresses resulting from the rotation loads relax the interference fit stress field sufficiently to cause
the positive radial deformations one would expect from a spinning rotor. The question now is
whether this models what would actually occur. This author believes that it is a possibility. If
one induces a large stress field into a material that experiences creep at a certain temperature and
then proceeds to place this material in that temperature environment, it is reasonable to expect
that material to experience creep deformations as a result of the induced stress field.

79
Figure 3-51 – Pole-tip Radial displacement vs. RPM for T=780K, Type A, d=0.0015 and d=0.003’’’’,
ATi=AfeCo

80
The radial displacement profile, strain fringe plot, and time history of radial displacements plots
for the 780K, d equal to 0.003 inch, and ω equal to 35,000 rpm case are found in Figures 3-52
through 3-54 respectively.

All of the analyses described in this section were also performed using the material variables
from the creep type AB1 model. The pole-tip radial displacements at various rotation speeds for
the interference fit equal to 0.0015 and 0.003 inch models can be seen in Figure 3-55. It was
previously mentioned that at a temperature around 780K it was expected that differences may
present themselves between the creep types A and AB1. A comparison of the results indicates a
very slight difference. It appears that any significant difference between the creep types would
occur at still greater temperatures, but since the survivable operating environment is clearly
getting further away from that which the Air Force outlined, this was not examined. The stress
and radial displacements profiles for d equal to 0.0015 inches at various rotation speeds are seen
in Figures C-9 through C-11, and the profiles for d equal to 0.003 inches are seen in Figures C-
12 through C-14.

3.2.2 Single Lamina Results – Radial Variation in Temperature Models

The results presented in section 3.2.1 all involved constant temperature rotors. In actuality the
shaft-rotor system would exhibit a temperature profile dependent on radial position. The
temperature of the “cooling” air exiting the compressor of an aircraft’s jet engine may be as high
as 1000°F (≈810K). The heat of the environment would then be transferred through the shaft-
rotor system and would result in a temperature profile where the rotor poles would be hotter than
the shaft. Since this report seeks to model steady-state creep the assumption is made that the
temperature profile has reached a steady-state (no heat transfer). It was decided that a linear
temperature profile with respect to radius would be used to model the temperature field. The
temperature range of 600 to 1000°F (588 to 810K) was chosen based on [51]. Two other
temperature ranges were chosen. The first was 850 to 1000°F (727 to 810K). This was
examined once it was realized that the first model would not exhibit significant creep
deformations. The third model was the reverse of the

81
Figure 3-52 – Shaft-Hub Radial displacement Profile for d=0.003’’, ω=35,000rpm, T=780K, t=642hrs, Type
A, ATi=AFeCo

ω=35,000rpm, T=780K, t=642hrs, Type A,


Figure 3-53 – Von Mises Strain Fringe Plot for d=0.003’’,ω
ATi=AFeCo

82
Figure 3-54 – Radial displacement vs. Time for d=0.003’’, T=780K, ω=35,000rpms, Type A, ATi=AFeCo

Figure 3-55 – Pole-tip Radial displacement vs. RPM for T=780K, Type AB1, ATi=AFeCo

83
first where the hot portion of the shaft-rotor system would now be at the shaft and the cooler
portion at the rotor pole-tips. The temperature range was the same as used in the first model.

3.2.2.1 588K – 810K

A rotation speed range of 45,000 to 55,000 rpm was examined for this temperature profile. The
results indicate that there is no failure prior to the 5,000 hr time limit for rotation speeds between
45,000 and 50,000 rpm. The results for the 55,000 and 53,000 rpm case both indicate an
interference fit failure at time zero which implies that failure occurs immediately upon loading,
see Figures 3-56 and 3-57. Therefore, between 50,000 and 53,000 rpm lies the maximum
rotation speed at which the shaft-rotor system can survive for 5,000 hrs.

The results indicate that the shaft-rotor system is not experiencing significant creep deformations
at the rotation speed of 50,000 rpm, see Figure 3-58. The slopes of the shaft and rotor hole lines
appear to be almost zero and the pole-tip radial displacement line has only a slight slope. The
plot in Figure 3-59 indicates that the stress profile does not noticeably change from 0.257
seconds to 5,000 hrs indicating that only elastic deformations are experienced prior to 5,000 hrs.
These results are due to the “threshold” nature of the creep properties reported in [3]. This
“threshold” phenomenon is most clearly depicted in Figure 3-3. This figure indicates a drastic
shift in the stress coefficient, which occurs at about 780K (850°F). At temperatures below 780K
the stress coefficient is essentially zero, therefore the strain rate in equation 2.2-1 becomes zero
as well, indicating that the material is not experiencing creep. The temperature profile being
modeled has a large interior portion of the rotor as well as the shaft at temperatures near or below
this “threshold” temperature. Therefore, a large portion of the system is not experiencing creep
deformations and would act to damp (or restrict) the creep deformations of the “hot” outer
portions of the rotor

The Air Force requirement that the shaft-rotor system be capable of operating in an environment
of 810K at a rotation speed of 55,000 rpm for 5,000 hrs now appears to be attainable with this
iron-cobalt alloy. These operating conditions were not possible for the constant temperature

84
Figure 3-56 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=588K, Tout=810K11, ω=53,000rpm, Type A

Figure 3-57 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=588K, Tout=810K, ω=55,000rpm, Type A

11
Tin is the temperature at the rotor hole and Tout is the temperature at the pole tips.

85
Figure 3-58 – Radial displacement vs. Time for d=0.0015’’, Tin=588K, Tout=810K, ω=50,000rpms, Type A

Figure 3-59 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=588K, Tout=810K, ω=50,000rpm, Type A

86
models of section 3.2.1. These conditions, however, may be attainable by the variable
temperature model presented in this section. The model presented above was subjected to a
temperature environment of 810K and it failed prior to 5,000 hrs at a rotation speed between
50,000 and 53,000 rpm, only a few thousand rpm less than required. Since the model does not
experience creep deformations the interference fit need only to be made strong enough to
withstand the elastic stress field resulting from the rotation load. Therefore, based on data
collected from the constant temperature models it is reasonable to assume that an increase in the
interference fit (d) would allow the shaft-rotor system to survive the environment stipulated by
the Air Force.

Results for the 45,000, 47,000, and 48,500 rpm cases are shown in Figures B-16 through B-18.
The 45,000 rpm case is presented in Figure B-16. This set of data indicates that little to no creep
has occurred in the 5,000 hrs. The radial stress decreases from zero at the shaft hole to about –
2000 psi at the rotor hole indicating that the interference fit has not been compromised. The
radial stress then increases to a maximum value of about 19,000 psi at about 1 inch. The
tangential stress in the shaft is at an almost constant –10,000 psi. This value jumps to about
58,000 psi at the rotor hole and decreases to an average value of about 20,000 psi at the hub
radius. The radial displacement vs. time plot, in Figure B-16, supports the conclusion that creep
is not effecting the overall deformations as indicated by the zero or near zero slopes of the
various curves. The 47,000 rpm results are presented in Figure B-17. The radial stress in the
shaft decreases from zero to about –1000 psi, or 1000 psi greater than the 45,000 rpm case. The
radial stress then increases to about 20,000 psi at 1 inch again about 1000-psi greater than the
25,000 rpm case. The radial displacement vs. time plot in Figure B-17 again shows that there is
no creep as indicated by the zero slope lines at the various radial positions. Results for the 48,500
rpm case are presented in Figure B-18. The increase in rotation speed does not result in creep
and there is only a slight difference in the stress and radial displacement profiles.

87
3.2.2.2 727K – 810K

The rotation speed range examined for this temperature profile was between 30,000 to 45,000
rpm. The results indicate that the shaft-rotor system does not experience failure between 30,000
and 40,000 rpm. Figure 3-60 shows that interference fit failure occurs between 664 and 3,786
hrs at 45,000 rpm. This hotter temperature profile experiences failure at a slower rotation speed
than the previous model, as one would expect. The plot in Figure 3-61 shows that this
temperature profile also does not experience a large amount of creep deformation. The radial
displacement curves do not have large slopes prior to failure for the 45,000 rpm case. However,
the radial displacement curves do have a higher slope than seen in the previous temperature
profile, indicating that the hotter profile is experiencing a greater amount of creep deformation,
see Figures B-19 through B-21. The 30,000 rpm results are presented in Figure B-19. The data
indicate a small amount of creep deformations with the small slopes in the radial displacement
vs. time plot. The radial stress decreases from zero at the shaft hole to about –8,000 psi at the
rotor hole followed by an increase to about 7,000 psi at 1 inch and a final decrease to an average
of about 4,000 psi at the hub radius. The radial stress profile at 5,000 hrs does not indicate a
significant change from the 0.257 seconds case. The tangential stress at 0.257-seconds starts at –
30,000 psi at the shaft hole and increases to about –28,000 psi at the rotor hole. There is a jump
to about 31,000 psi, which then falls to an average value of about 10,000 psi at the hub radius.
An increase to 35,000 rpm is presented in Figure B-20. The pole-tip radial displacement has
increased from about 8.4 × 10 −4 inches for the 30,000 rpm case to about 1.22 × 10 −3 inches at
35,000 rpm. The radial stress falls to about –8,000 psi at the rotor hole and then increases to
about 10,000 psi at 1 inch. The maximum tangential stress decreases from 39,000 psi at the rotor
hole to an average of about 12,000 psi at the hub radius. The 40,000 rpm results are presented in
Figure B-21. The radial stress in the shaft indicates that the shaft-rotor system is near failure.
The radial stress at the rotor hole is about –5,000 psi at 0.257 seconds, which increases to about –
2000 psi at 5,000 hrs. The tangential stress in the rotor falls from 47,000 psi to an average of
about 16,000 psi at the hub radius.

88
Figure 3-60 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=727K, Tout=810K, ω=45,000rpm, Type A

Figure 3-61 – Radial displacement vs. Time for d=0.0015’’, Tin=727K, Tout=810K, ω=45,000rpms, Type A

89
This hotter temperature profile results in a decrease in survivable rotation speed as was seen in
the constant temperature data, see Figure 3-62. The cooler temperature profile of 588K to 810K
experiences failure between 50,000 and 53,000 rpm while the hotter profile of 727K to 810K
fails between 40,000 and 45,000rpm. This behavior reflects the “threshold” nature of the creep
model and material properties reported in [3]. Failure is the result of the rotor becoming
separated from the shaft. This occurs at slower rotation speeds for higher temperatures because
at these temperatures the material properties indicate a greater creep rate. The pole-tip
deformations are still well below those determined to cause failure from pole-tip rubbing. Even
at the higher temperatures the creep rate is not large enough for the creep deformations to greatly
exceed the magnitude of the elastic deformations.

If one compares the results for this temperature profile to the constant temperature profile of
727K with d equal to 0.0015 inches, and ω equal to 40,000 rpm, one observes that the variable
temperature profile, as expected, experiences a greater amount of creep. This is because the
shaft-rotor system is basically hotter for the variable temperature case and, therefore, the creep
strain rate is greater than the cooler constant temperature model.

An examination of all the results presented thus far reveals that regardless of the temperature
profile the creep deformations experienced prior to failure are of the same magnitude as the
elastic deformations. According to Whal [30], if the creep strains are significantly greater than
the elastic strains then the transient period can be neglected. It is clear form this analysis that the
creep strains are not large enough, based on Whal, to ignore the effects of the transient period.
Wahl indicates that the transient period yields deformations of the same magnitude as the elastic
deformations. Therefore, the actual deformations should be about twice that of those presented
in this paper. These new deformations would still be significantly less than that required to
cause pole-tip rubbing; therefore, the results published in this report are a good indicator of the
survivable conditions for the shaft-rotor system. The transient conditions should be taken into
account to obtain a more accurate representation of the radial displacement and stress field.

90
Figure 3-62 – Pole-tip Radial displacement vs. RPM for Three Variable Temperature Profiles, Type A,
d=0.0015’’

91
3.2.2.3 810K – 588K

The rotation speed range of 25,000 to 35,000 rpm was examined for this temperature profile.
The results indicate that there was no failure prior to the 5,000 hrs time limit for rotation speeds
between 25,000 and 30,000 rpm. The results for the 35,000 rpm case indicate an interference fit
failure between 2,008 and 5,000 hrs, see Figure 3-63. The radial displacement time history
reflects the temperature profile of this model. The shaft and rotor hole lines show a higher creep
rate (slope) than the pole-tip line. The hotter inner portion of the shaft-rotor system is
experiencing a higher creep rate than the outer portion, see Figure 3-64. The radial displacement
profile does not reflect the general trend that has been seen in most of the previous results. The
large creep rate of the shaft is more clearly seen in Figure 3-65 where the radial displacement
after 2,008 hrs shows that the shaft has displaced further in magnitude than the inner portion of
the rotor, which is at a cooler temperature. As the radial position moves toward the outer poles,
it is apparent that the outer portions are restricting the radial displacement of the inner portion, as
apparent in the compression represented by the dip in magnitude of the radial displacements.
The radial displacements then increase as the radial position increases, revealing that these outer
sections are being pushed by the cumulative radial displacements of the inner portions.

The temperature profile presented in this section fails at the lowest rotation speed of the three
variable temperature profiles studied, see Figure 3-62. Failure occurs between 30,000 and
35,000 rpm as opposed to between 40,000 and 45,000 rpm for the 727K-810K profile, and
between 50,000 and 53,000 rpm for the 588K-810K profile. The temperature profile plays a
significant role in the deformation profile. The 588K-810K profile does not experience creep
radial displacements and can survive at rotation speed up to 15,000 rpm higher than the 810K-
588K profile. The profile with the higher inner radius temperature (810K-588K) experiences
small amounts of creep deformations but cannot survive at the higher rotation speeds important
to this study.

92
Figure 3-63 - Shaft-Hub Stress Profile for d=0.0015’’, Tin=810K, Tout=588K, ω=35,000rpm, Type A

Figure 3-64 – Radial displacement vs. Time for d=0.0015’’, Tin=810K, Tout=588K, ω=35,000rpms, Type A

93
Figure 3-65 – Shaft-Hub Radial displacement Profile for d=0.0015’’, ω=35,000rpm, Tin=810K, Tout=588K,
t=2008hrs, Type A

94
The initial decrease in radial displacements seen in the constant temperature model, Figures B-14
and B-15, can also be seen in the lower rotation speeds of this temperature profile, see Figures B-
22 and B-23. This trend indicates that initially the interference fit stress filed is dictating the
creep deformations. This radial displacement profile is not seen in the 30,000 rpm case for this
temperature profile, see Figure B-24. At a rotation speed between 27,000 and 30,000 rpm the
stresses induced by the rotation load become large enough to cause creep radial displacements in
the positive radial direction.

The radial stress for the 25,000 rpm case presented in Figure B-22, increases from about –9,000
psi at 0.257 seconds to about –1000 psi at 5,000 hrs. There is a large shift in the tangential stress
at the shaft hole from –40,000 to –5,000 psi between 0.257 seconds and 5,000 hrs. The
tangential stress in the rotor follows a new profile at 5,000 hrs. It begins at about 6,000 psi and
then increases to a maximum value of about 12,000 psi around 0.875 inches. The tangential
stress finally decreases to an average value of about 9,000 psi at the hub radius. The radial
displacement vs. time plot shows that after the initial positive elastic radial displacement the
creep deformations are in the negative (radial) direction before eventually reversing to a positive
(radial) radial displacement. After 5,000 hrs of creep the final pole-tip radial displacements are
slightly less than the initial elastic radial displacement. The results for the 27,000 rpm case are
presented in Figure B-23. The radial displacement and stress profiles are similar to the results in
Figure B-22. The radial displacement vs. time plot indicates that the initial negative (radial)
creep displacements are not as dominant as in the previous case. This can be seen by the fact
that the final pole-tip radial displacement exceeds the initial elastic radial displacement.

3.2.3 Multiple Lamina Stack Results - Axial Variation in Temperature Models

Sections 3.2.1 and 3.2.2 both involve models that examine the effects of the operating
environment on a single lamina of the rotor stack. The rotor would consist of hundreds of lamina
compressed together by an axial force such as that represented in Figure 2-4. Therefore, in order
to obtain a more realistic model of the shaft-rotor system, multiple lamina stacks need to be

95
Figure 3-66 – Two Lamina FE Mesh

96
examined. The multiple lamina models would need to account for the surface interactions
between the lamina as well as the effects of an axial as well as radial variation in temperature.
This section contains results that represent a first step towards the development of this more
complete multiple lamina model.

The FE models in this section involve two laminas held together by a compressive axial force.
Both lamina are modeled with a constant temperature profile (similar to section 3.2.1) though
each lamina is at a different temperature, thereby simulating an axial temperature variation.
These models do not contain the shaft nor the initial stress field induced by the interference fit.
In an attempt to bound the radial displacement results each model is examined with the inner
radius of the rotor “clamped” and “unclamped”, where the “clamped” model simulates a rotor
attached to a rigid body at the inner radius of the rotor, and the “unclamped” model is free to
deform at the inner radius of the rotor. In actuality the shaft and induced stresses would effect
the resulting stress and displacement fields in such a way that the final displacement field would
fall between the clamped and unclamped cases. The FE model is comprised of 896 elements, as
shown in figure 3-66. This FE model contains fewer elements than the single lamina models but
required up to fifteen times as much wall clock time to converge to a solution due to the surface
interaction calculations.

The surface interaction between the lamina was not known at the time of this report. Therefore,
a simple Coloumb (static) friction model with finite sliding was examined. Each model was
examined for two values of Coloumb friction. The first value was zero, or a frictionless sliding
case. The second value was 0.5, which indicates that if the forces transmitted tangentially
between the two surfaces exceed 50% of the axial forces then slipping will occur.

Finally the rotation speeds examined were 45,000 and 55,000 rpm.

Each lamina is 0.12 inches thick or 20 times the thickness of the single lamina models. The
entire stack is assumed to be 1.2 inches thick. This would represent 200 laminae of 0.006 inches
thick or 10 laminae of 0.12 inches thick. The laminas were modeled 20 times thicker so that
there would be a greater temperature difference in the axial direction. The temperature profile is

97
assumed to be linear with the Z equal to 0 inch surface of the bottom rotor is assumed to be 810
K and the Z equal to 1.2 inch surface of the last lamina is 588 K. The temperature of the lamina
was determined by calculating the temperature at the center (in the Z or thickness direction) of
the lamina. The temperature of the bottom lamina is 798.9 K and the top or second lamina is
794.7K. Even with the lamina models 20 times thicker it is obvious that there is very little
temperature variation between each lamina. The temperature variation would be almost
negligible if the lamina were modeled actual size.

3.2.3.1 The 45,000 rpm Models

The rotation speed equal to 45,000 rpm models are presented in this section. The first model is
for a frictionless sliding (or µ = 0) model with an unclamped inner rotor (or hub) radius. The
time histories of displacements at the inner hub radius and pole tips can be seen in Figure 3-67.
It is clear that the “hotter” bottom lamina is experiencing more creep than the top lamina. This is
indicated by the greater final displacements of the inner hub radius and pole tips. The bottom
lamina has displaced about 3.8 × 10 −2 inches at the inner hub radius and about 3.5 × 10 −2 inches at
the pole tips while the top lamina displaces about 2.8 × 10 −2 and 2.5 × 10 −2 inches at the hub
radius and pole tips respectively. These values also indicate that the outer portions of the rotor
are confining the displacements of the inner portions as can be seen from the greater hub
displacements. The result is a compression of the rotor or a decrease in the distance from the
hub radius to the pole tips. A three dimensional (3D) view of the displacements is shown in
Figure 3-68. Again the difference in displacements as well as the lack of friction can be seen in
this figure. The inner portions of the rotor experience a larger amount of axial compression than
the outer portion of the rotor, specifically the rotor poles. Comparing the undeformed outline to
the deformation plot in Figure 3-68 shows this. The reason for this lies in the fact that the inner
portions of the rotor experience a greater amount of stress, as seen in Figure 3-69 and 3-70. The
higher stress is indicative of a larger displacement which “stretches” the lamina more in these
areas resulting a reduction in thickness. The pole tips experience a small amount of stress and as
a result there is no apparent reduction in the thickness (in the Z direction) of the poles.

98
Figure 3-67 – Displacement vs. Time for 2 Lamina Model, Inner Radius unclamped, ω=45,000rpm, µ=0

99
Figure 3-68 – Deformation Plot at 5000 hrs for 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0

100
Figure 3-69 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius unclamped, , ω=45,000rpm, µ=0

101
Figure 3-70 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius unclamped, , ω=45,000rpm, µ=0

Figure 3-71 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius unclamped, , ω=45,000rpm, µ=0

102
The stress profiles, seen in Figure 3-69, clearly show the relaxation of both the radial and
tangential stress as a result of the creep deformations. The profiles at 5000 hrs for the top and
bottom lamina do not show a noticeable difference. Since the difference in temperatures
between the laminas is small it is expected that the actual difference in stress profile would be
small. The Von Misses stress profile, seen in Figure 3-70, shows that the inner portions of the
rotor are experiencing higher stresses than the pole tips. There are also high stress
concentrations where the poles meet the outer hub. Figure 3-71 shows the Von Misses strain
profile at 5000 hrs. As expected the highest strains are located along the inner radius of the hub
and the lowest strains are located in the poles. The pull of the poles can be seen in the strain
gradient lines. At a given radius the highest strains are found along the 45° lines. The material
along this line is experiencing the pull of both poles equally and in opposite directions resulting
in the higher stresses and strains.

The next case is identical to the previous case except that for this case the inner radius of the hub
is clamped (or prevented from deforming). Figure 3-72 shows the time history of the radial
displacement at the pole tips. At 5000 hrs the pole tip displacement for the bottom lamina is
about 9.5 × 10 −3 inches, or about 26% of the displacement experienced by the unclamped model.
The top lamina displaces about 7.0 × 10 −3 inches at the pole tips after 5000 hrs. This is 28% of
the displacement experienced by the unclamped model. It is clear that clamping the inner hub
radius reduces or damps the radial displacement of the pole tips. While the displacements have
been greatly reduced as a result of clamping the inner hub radius, the percent difference between
the bottom and top lamina for each case has not changed as much. The pole tip displacement for
the bottom lamina in the unclamped case is about 73.7% of that of the top lamina for the same
case. The pole tip displacement for the bottom lamina in the clamped case is about 71.4% of that
of the top lamina for the same case. This is expected since the temperature profile has not
changed. The difference is displacement profile can be seen in Figure 3-7312. Once again the
same

12
Keep in mind that all of the 3D deformation plots the displacements are not to scale. For viewing purposes the maximum displacement is
plotted as 0.1 times the model length which is a value determined by PATRAN (the program used to generate these plots).

103
Figure 3-72 – Displacement vs. Time 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0

Figure 3-73 – Deformation Plot at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0

104
general profile for axial displacement can be seen where the inner portions of the rotor
experience a greater amount of axial compression.

Figures 3-74 and 3-75 show the stress profiles and fringe plot for this case. Unlike the previous
case where the inner radius was free to deform and therefore the radial stress at the inner radius
is equal to zero, the radial stress for the clamped case is about 30,000 psi after 5000 hrs. The
tangential stress at the inner radius for the unclamped case fell from about 50,000 psi to about
35,000 psi over the 5000 hr lifetime. For the same time period the tangential stress for the
clamped case increases from about 12,000 psi to about 16,000 psi at the inner radius. The Von
Misses stress profile also shows the effect of clamping the inner radius. The maximum stress
and strain for a given radial position occurred at the 45° line for the unclamped case. This is not
the case for the clamped case. Clamping the inner radius appears to greatly constrain the stresses
and strains along the 45° line as compared to the unclamped case. The Von Misses strain plot is
shown in Figure 3-76.

The third 45,000 rpm case adds friction between the two laminae, the coefficient of friction (µ) is
equal to 0.5, and the inner radius for this case is again unclamped. Figure 3-77 shows the time
history of displacements. Comparison to Figure 3-67 clearly reveals the effect of the increased
coefficient of friction. The top and bottom laminae now basically experience the same
displacements as can be seen by the overlapping curves of the inner hub radius and pole tip
displacements for the two laminae. This indicates that the tangential forces between the two
surfaces are not significantly exceeding the axial forces between the two surfaces, therefore there
is little to no slipping. The result is that the frictional force forces the two laminae to displace in
the same manner, in a sense the “hotter” bottom lamina pulls the “cooler” top lamina along while
at the same time the top lamina retards the displacement of the bottom lamina. Comparing the
displacements at 5000 hrs between the µ equal to 0 and the µ equal to 0.5 cases this becomes
clearer. The pole tip displacement of the bottom lamina for the frictionless case is about
3.5 × 10 −2 inches, and it decreases to about 3.3 × 10 −2 inches in the friction case. The inner hub
radius for the bottom lamina displaces about 2.5 × 10 −2 inches in the frictionless case. This
displacement increases to about 3.0 × 10 −2 inches in the friction case. The results indicate that
the bottom lamina pulls with a greater force than the retarding force of the top lamina. Figure

105
Figure 3-74 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius clamped, , ω=45,000rpm, µ=0

106
Figure 3-75 - Von Misses Stress Fringe Plot 2 Lamina, Inner Radius clamped, , ω=45,000rpm, µ=0

Figure 3-76 - Von Misses Strain Fringe Plot 2 Lamina, Inner Radius clamped, , ω=45,000rpm, µ=0

107
Figure 3-77 – Displacement vs. Time for 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5

108
3-78 shows the 5000 hr displacement profile for the top and the bottom laminae. Again the
effect of the friction interface can be seen since the profile for the two laminae is basically
identical. Figure 3-79 shows the 3D deformation plot, again showing nearly identical radial
displacement profile of the two laminae.

The stress profiles and stress and strain plots can be seen in Figures 3-80 through 3-82
respectively. The radial and tangential stress profiles are very similar to the profiles for the
frictionless case both in magnitude and trend. There is a very small difference in the stress
profiles for the top lamina. These profiles no longer directly overlap those of the bottom lamina
as in the frictionless case. This difference is the result of the frictional force, which appears to
have a greater effect on the “cooler” top lamina as was seen in the displacement results. The
stress and strain fringe plots reveal similar trends to those in the frictionless case with small
differences in the magnitudes.

The final 45,000 rpm case models µ equal to 0.5 with the inner radius clamped. Again the
frictional force acts to retard the bottom lamina’s radial displacements while “pulling” or
increasing the displacements of the top lamina, see Figure 3-83. The bottom lamina’s pole tip
displacement at 5000 hrs for the frictionless case was about 9.5 × 10 −3 inches, which fell to about
8.2 × 10 −3 inches, or 86% of the frictionless value, for the friction case. The top lamina’s pole tip
displacement at 5000 hrs for the frictionless case was about 7.0 × 10 −3 inches, which increased to
about 8.0 × 10 −3 inches, or 114% of the frictionless value, for the friction case. It is interesting to
note that for this (friction) case the bottom lamina’s pole tip displacement is retarded almost the
same magnitude as the top lamina’s is increased. This was not the case for the unclamped
friction case where the top lamina’s pole tip displacement was increased by 120% over the
frictionless case, and the bottom lamina’s pole tip displacement was 94% of the frictionless case
at 5000 hrs. This suggests that the clamped inner hub radius increases the retarding force of the
“cooler” top lamina or conversely decreases the “pulling” force of the “hotter” bottom lamina.
The 3D deformation plot at 5000 hrs can be seen in Figure 3-84.

Figure 3-85 shows the radial and tangential stress profile. The stress profiles for the top and the
bottom laminae in the frictionless case, seen in Figure 3-74, appear to overlap each other

109
Figure 3-78 – Radial Displacement Profile for 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5

Figure 3-79 – Deformation Plot at 5000 hr for 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5

110
Figure 3-80– Radial and Tangential Stress Profile 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5

111
Figure 3-81 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5

Figure 3-82 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius unclamped, ω=45,000rpm, µ=0.5

112
Figure 3-83 - Displacement vs. Time for 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0.5

113
Figure 3-84 – Deformation Plot at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0.5

114
Figure 3-85 – Radial and Tangential Stress Profile 2 Lamina, Inner Radisu clamped, ω=45,000rpm, µ=0.5

115
at 5000 hrs. This is not the case for the friction case where the top lamina’s stress profiles are
slightly higher in magnitude than the bottom lamina’s. Further comparison to the frictionless
case reveals that the bottom lamina’s stress profile is lower for the friction case while the top
lamina’s stress profile is a little greater in magnitude for the friction case. The stress and strain
fringe plots are in Figures 3-86 and 3-87.

While the two friction cases presented thus far do not exactly predict the “real world”
displacements it can be concluded that the clamped and unclamped cases bound a more realistic
displacement value. The only failure criterion applicable to these cases is pole tip rubbing as
defined previously in this report. The pole tip displacement after 5000 hrs would most likely fall
between 8.0 × 10 −3 and 3.0 × 10 −2 inches, see Figures 3-77 and 3-83. The possible displacements
do not bound the failure criterion of 0.24 to 0.4 inches, therefore pole tip failure does not occur
in these models.

3.2.3.2 The 55,000 rpm Models

The rotation speed equal to 55,000 rpm cases are presented in this section. The first case present
is the frictionless unclamped model. Increasing the rotation speed has the predictable effect of
increasing the magnitude of the final displacements. Pole tip displacement after 5000 hrs for the
45,000 rpm case was about 3.5 × 10 −2 and 2.5 × 10 −2 inches for the bottom and the top laminae
respectively, see Figure 3-67. These displacements increased 429 and 440% to 1.5 × 10 −1 and
1.1 × 10 −1 inches with the increase of the rotation speed, see Figure 3-88. The 5000 hr
displacement profiles can be seen in Figure 3-89. The difference in the displacement profile
between the two laminae is clearly depicted. The top or “hotter” lamina has displaced more than
the “cooler” top lamina, as was seen in the frictionless 45,000 rpm case.

Figure 3-90 shows the radial and tangential stress profiles. As was seen in the 45,000 rpm case
the 5000 hr stress profiles for the top and the bottom laminae apparently overlap for the
frictionless case. This is due to the small temperature difference between the two laminae, which
seems to have little to no impact on the stress profiles. The basic stress profile trends for the two

116
Figure 3-86 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0.5

Figure 3-87 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius clamped, ω=45,000rpm, µ=0.5

117
Figure 3-88 – Displacement vs. Time for 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0

118
Figure 3-89 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0

119
Figure 3-90 – Radial and tangential Stress Profile 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0

120
unclamped frictionless cases are the same with only the magnitudes changing with the increase
in the rotation speed. The maximum radial stress for both rotation speed cases occur at about 1
inch. The elastic (0.26 SEC) radial stress value is about 17,500 psi for the 45,000 rpm case,
which then increases to about 27,000 psi for the 55,000 rpm case. Similarly the maximum
tangential stress, which occurs at the inner hub radius, increases from about 50,000 to 75,000 psi
between the 45,000 and 55,000 rpm case. Figures 3-91 and 3-92 show the stress and strain
fringe plots respectively. These plots are identical to the 45,000 rpm case in all but the actual
values. As expected the increase in the rotation speed results in an increase in the stress and
strain values.

The next set of results are from the frictionless clamped case. Once again a large increase in
displacement is seen as a result of the increase in rotation speed. The pole tip displacement has
increased about 393% from about 7 × 10 −3 to about 2.75 × 10 −2 inches for the top lamina, see
Figures 3-72 and 3-93. The pole tip displacement for the bottom lamina increases about 400%
from about 9.5 × 10 −3 to about 3.8 × 10 −2 inches. Figure 3-94 shows the displacement profile.
Similar to Figure 3-89, the difference in displacement profile can be seen in this frictionless case,
though the two laminae’s displacement profiles overlap more than in the unclamped case.

The radial and tangential stress profiles are shown in Figure 3-95. These profiles depict the same
trend as seen in the similar 45,000 rpm case. The maximum value of the radial stress occurs at
the inner hub radius, which increases from about 28,000 to about 38,000 psi at 5000 hrs for both
the top and the bottom laminae between the 45,000 and 55,000 rpm cases. Similarly, the
maximum tangential stress, which occurs at about 1 inch, increases from about 21,500 to about
33,000 psi for both the top and the bottom laminae between the two rotation speed cases. The
stress and strain fringe plots are shown in Figures 3-96 and 3-97 respectively.

The next 55,000 rpm case presented is the unclamped µ equal to 0.5 model. Figure 3-98 shows
the displacement time history. The pole tip displacement increased about 433% from about
3.0 × 10 −2 to about 1.3 × 10 −1 from the 45,000 to the 55,000 rpm case at 5000 hrs. Comparing the
unclamped 55,000 rpm case to the clamped 55,000 rpm case the effect of the friction interaction
is seen to be similar to that observed in the 45,000 rpm cases. The top or “cooler”

121
Figure 3-91 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0

Figure 3-92 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0

122
Figure 3-93 – Displacement vs. Time for 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0

Figure 3-94 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0

123
Figure 3-95 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0

124
Figure 3-96 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0

Figure 3-97 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0

125
Figure 3-98 – Displacement vs. Time for 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5

126
lamina 5000 hr pole tip displacement increases from about 1.1 × 10 −1 to about 1.3 × 10 −1 inches
while the bottom lamina pole tip displacement decreases from about 1.5 × 10 −1 to about 1.3 × 10 −1
inches. It appears that since the final pole tip displacements for the top and bottom laminas are
equal for the 55,000 rpm friction case that the tangential stresses do not exceed 50% of the axial
stresses between the two surfaces which indicates that no slipping has occurred. This can be
clearly seen in the displacement profile in Figure 3-99.

The radial and tangential stress profiles are shown in Figure 3-100. The maximum radial stress
is at about 1 inch as was seen in the 45,000 rpm case, see Figure 3-80 . The maximum radial
stress at 0.26 seconds increases from about 17,000 to about 26,000 psi between the 45,000 and
the 55,000 rpm cases. Similarly, the maximum tangential stress, which occurs at the inner hub
radius, increases from about 50,000 to about 75,000 psi between the 45,000 and the 55,000 rpm
cases respectively. The stress and strain fringe plots are shown in Figure 3-101 and 3-102. They
show the same trends with different values as seen in the 45,000 rpm case.

The final model presented in this section is the 55,000 rpm inner radius clamped and µ equal to
0.5 case. Clamping the inner hub radius reduces the 5000 hr pole tip displacement as was seen in
the 45,000 rpm cases, see Figure 3-103. The pole tip displacement decreases from about
1.3 × 10 −1 to about 3.3 × 10 −2 inches between the unclamped and the clamped 55,000 rpm cases.
Comparing to the same 45,000 rpm case seen in Figure 3-83 to this case the effect of the increase
in the rotation speed can again be seen.. The 5000 hr pole tip displacement increases from about
8.3 × 10 −3 to about 3.3 × 10 −2 inches between the 45,000 and the 55,000 rpm cases respectively.
Figure 3-104 shows the displacement profile for this case. It is clear that there is little to no
slipping occurring since the top and the bottom laminae exhibit overlapping displacement
profiles.

The stress profiles are shown in Figure 3-105. Once again, the profiles are similar to the 45,000
rpm case only the values have changed. The maximum radial stress has increased from about
30,000 to about 45,000 rpm between the 45,000 and the 55,000 rpm cases at the inner hub radius
for the top lamina at 5000 hrs. Similarly the maximum tangential stress increases from about
23,000 to about 34,000 psi between the 45,000 and the 55,000 rpm cases at 1 inch for the top

127
Figure 3-99 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5

128
Figure 3-100 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5

129
Figure 3-101 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5

Figure 3-102 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius unclamped, ω=55,000rpm, µ=0.5

130
Figure 3-103 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5

Figure 3-104 – Displacement Profile at 5000 hrs for 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5

131
Figure 3-105 – Radial and Tangential Stress Profile 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5

132
lamina at 5000 hrs. Once again the 5000 hr stress profiles show a slight difference between the
top and the bottom laminae. This is the effect of clamping the inner radius since this was not
seen in the previous unclamped case. The stress and strain fringe plots are shown in Figures 3-
106 and 3-107 respectively.

Once again the friction clamped and unclamped results can be considered to bound a more
reasonable displacement profile. Therefore, the pole tip displacement after 5000 hrs would fall
between 3.3 × 10 −2 and 1.3 × 10 −1 inches, which does not indicate pole tip rubbing has occurred.
These models presented in section 3.2.3 represent a simplified look at the interaction between
two laminae. The interaction between the laminae need to be further studied in order to provide
a better representation for this interface. Models containing more laminae would also be a
logical next step in this research. Until such a time as the multiple laminae models are studied in
greater detail it is this author’s opinion that the radial variation in temperature single lamina
models provide the most useful information.

133
Figure 3-106 – Von Misses Stress Fringe Plot 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5

Figure 3-107 – Von Misses Strain Fringe Plot 2 Lamina, Inner Radius clamped, ω=55,000rpm, µ=0.5

134
4 Conclusions/Summary

4.1 Accomplishments

A large number of finite element (FE) models simulating various operating environments are
presented in this report and contribute to the body of knowledge pertaining to the use of iron-
cobalt alloys in the IPU of a MEA. The elastic results presented provide a baseline with which
to compare the creep results. Constant temperature models representing worst case scenarios
were studied and provided “safe” radial displacement and stress profiles. Next, more realistic
radial variation in temperature models were studied using a linear variation in temperature with
respect to radial position. These results indicate an operational range without a built in safety
factor. Finally, a few variable temperature with respect to axial position models were studied to
expand the research to multiple laminae. The lamina stack was modeled with the Coulomb
friction between the adjoining rotor laminae.

4.1.1 The Finite Element Model

A program created by this author generated the FE models used in this report. The program is
capable of generating a complete ABAQUS input file for various rotor and shaft dimensions and
operating conditions. This program generates input files for many combinations of the shaft-
rotor system such as a shaft tied to a rotor, a shaft tied to a disk (rotor without poles), a rotor with
unclamped inner radius, a rotor with clamped inner surface, and a rotor lamina stack with
Coulomb friction between the laminae. The input files are generated by inputting the rotor
dimensions and the operating environment variables which are read by the program genmesh.F.
The user’s manual for this program and the post processing code post.F are given in Appendix A.

135
4.1.2 Creep Results

Three types of creep results are presented in this report. The first is the set of constant
temperature single lamina models. These provide worst case scenarios since the entire shaft and
the rotor are assumed to be at the same temperature. This assumption results in lower or safer
operating temperatures than the more realistic radial variation in temperature models. The
constant temperature models cannot survive for 5000 hrs at the required 55,000 rpm rotation
speed and 1000°F environment set forth by the Air Force. The constant temperature models
enable one to conclude that the interference fit failure occurs prior to the pole-tip rubbing which
was indicated in the elastic analysis. The effects of the interference fit on the survivable
operating conditions can also be seen in this analysis.

The radial variation in temperature models provide a more realistic indication of the survivable
operating environment. The cooler interior of the rotor damps or restricts the hotter outer
portions which are experiencing a greater magnitude of creep. The result of this profile is that
the shaft-rotor system does not experience a significant amount of creep and therefore the elastic
deformations dominate the overall deformations. According to Wahl [30], when the creep
strains are not significantly greater than the elastic strains, the primary stage of creep can no
longer be ignored. The primary creep deformations are usually on the order of the elastic
deformations and therefore are only important when the steady state creep deformations do not
greatly exceed the elastic deformations. The deformations presented in this report therefore may
not represent the total deformation profile. If the deformations from primary creep are of the
same magnitude as the elastic deformations, it is obvious that the shaft-rotor system would still
not fail as a result of the pole-tip rubbing. The increase in radial displacements results in a shift
in magnitude of the stress profile and may cause interference fit failure at a time sooner than
those presented in this report.

The last set of data presented in this report involves multiple laminas with an axial variation in
temperature. The temperature profile involves constant temperature rotor laminas stacked on top
of each other and held together by a compressive axial load. The surface interface between the

136
laminas is assumed to be the Coulomb or static friction. These results, like those of the single
lamina constant temperature models, represent a worse case scenario since the more realistic
temperature profile would involve a variation in the radial and axial directions. This data is a
good first step towards understanding the behavior of the entire lamina stack. A more realistic
friction model should be determined in order to attain more confidence in the results.

4.2 Applications

The codes developed for this study can be used with the FE analysis program ABAQUS to
quickly generate radial displacement and stress profiles for various rotor geometries and
operating environments. The code uses the power law creep model but could be easily edited to
include other creep models included in ABAQUS if necessary. The modeling of the entire shaft-
rotor system would necessitate a FE model so large that solution times would likely be on the
order of days or weeks rather than the mere hours necessary for the models presented here.

The FE model was set up so that the shaft was “tied” to the rotor. This is accomplished by
restricting the radial displacement of the nodes on the shaft outer radius to be equal to those of
nodes on the rotor hole. This allows the FEA to continue past the interference fit failure, the
point where the rotor becomes separated from the shaft. The end result is that the analysis does
not stop once failure occurs and thereby cannot easily be used to determine the time of failure for
those cases that fail prior to the 5,000 hr limit. This is why failure is indicated by a range of
times in the result section of this report. Since the focus of this report was to determine the
operating environment where the shaft-rotor system survives for 5,000 hrs, it was decided that
knowing the exact time of failure was not necessary.

Results from the radial variation in temperature models provide the most representative set of
operating conditions. The temperature is assumed to vary linearly with respect to radius. This
variation was chosen because of the simplicity of the resulting temperature profile. More
accurate profiles can easily be implemented in the mesh generating code if they become

137
available. Until these profiles are determined the ones presented in this report are sufficient to
provide useful indications of the survivable operating environments.

4.3 Future Research

This report used the material model presented in [3] to determine the safe operating environment
for the shaft-rotor system. Therefore, the results are only as good as the material model used.
The only way to verify the validity of this model is to perform laboratory spin tests to confirm
the results presented in this report. Until such tests are preformed it cannot be concluded with
any certainty whether or not these models accurately represent the “real world” behavior of the
shaft-rotor system. Until such a time as the laboratory tests are performed these models should
only be used as a preliminary, not definitive, design tool.

The strain model presented in this study involves the effect of the elastic and steady state creep
deformations. This model does not include the effects of the initial transient period nor the
effects of thermal expansion of the materials. While the transient period may not have a
significant effect on the results it is conceivable that thermal expansion could have a significant
effect. If the material of the shaft is chosen such that the coefficient of thermal expansion (CTe)
would result in the shaft deforming at a higher rate than the iron-cobalt alloy rotor the
interference fit would remain intact longer than seen in the models presented in this study. It
would be good to research weather or not the thermal expansion of the materials could delay or
even negate interference fit failure.

The results presented in this report reflect a solely mechanical analysis. The iron cobalt-alloy
was chosen not only for its mechanical properties, but also for its magnetic and electrical
properties. To date these three properties have been analyzed separately. The next step in
computer modeling should incorporate these three properties and their interaction. To the
author’s knowledge no such attempt to combine a mechanical, magnetic, and electric analysis
has ever been done.

138
5 References

1. Quigley, R. E., “More Electric Aircraft,” Conference Record, IEEE Applied Power
Electronics Conference, March 7-11, pp. 906-911, 1993.
2. Wiemer, J., A., “Electrical Power Technology For The More Electric Aircraft,” Conference
Record, AIAA/AHS/ASEE Aerospace Design Conference, February 16-19, p. 445-450,
1993.
3. Fingers, Richard T., “Creep Behavior of Thin Laminas of Iron-Cobalt Alloys for Use in
Switched Reluctance Motors and Generators,” Ph.D. Dissertation, Virginia Polytechnic
Institute and State University, June 17, 1998.
4. Wiemer, Jospeh A., “Power Management and Distribution for the More Electric Aircraft”,
Aerospace Power System and Technology Proceedings of the Intersociety Energy
Conversion Engineering Conference, Vol. 1, 1995, p.273-277.
5. Fingers, R.. T., “Air Force Application of Advanced Magnetic Materials”, Materials
Research Society Symposium Proceedings, Vol. 577, 1999, p.481-486.
6. Nippes, P. I., “Magnetism and Stray Currents in Rotating Machinery,” Technical Brief, Vol.
118, Journal of Engineering for Gas Turbines and Power, pp. 225-228, January 1996.
7. Miller, T. J. E., “Faults and Unbalance Forces in the Switched Reluctance Machine,” IAS
1993 Conference Record, Vol. 1, p. 87-96, October 2-8, 1993.
8. Dowling, Norman E., Mechanical Behavior of Materials, Prentice Hall, Upper Saddle River,
New Jersy, 1999.
9. Greenfield, P., Creep of Metals at High Temperatures, Mills and Boon Ltd., London, pp.7-
25, 1972.
10. Andrade, E. N. DaC., “On the Viscous Flow of Metals, and Allied Phenomena,”
Proceedings of the Royal Society, Section A, 84, pp. 1-12, 1920.
11. Andrade, E. N. DaC., “The Flow in Metals Under Large Constant Stresses,” Proceedings of
the Royal Society, Section A, Vol. 90, pp.329-341,1914.
12. Cadek, Josef, Creep in Metallic Materials, Elsevier, Amsterdam, pp. 60-69, 205-210, 246-
249, 1988.
13. Burton, B., Diffusional Creep of Polycrystalline Materials, Trans Tech Publications, Bay
Village, Ohio, pp.1-7,52-56, 1977.
139
14. Nabarro, F. R. N., H. L. de Villiers, The Physics of Creep, Taylor and Fracis Ltd., Bristol,
PA, pp. 1-2, 15-22, 32-36, 47-49,53-78, 1995.
15. Nabarro, F. R. N., “Report of a Conference on the Strength of Solids,” Physical Society,
London, p.75, 1948.
16. Weertman, J., “Steady-State Creep of Crystals,” Journal of Applied Physics, 28, pp.1185-
1189, 1957.
17. Raj, R. and M.F. Ashby, “On Grain Boundary Sliding and Diffusional Creep,” Metallurgical
Transactions, Vol. 2, pp-1113-1127, 1971.
18. Söderquist, Bertil A. T., Creep Under Equal Biaxial Loading of a Disk With a Hole, Acta
Polytechnica Scandinavica Physics Including Nucleonics Series No. 51,Stockholm, 1968.
19. Laks, H., Effect of Stress on Creep at High Temperatures, Berkeley, Minerals Research
Laboratory, Institute of Engineering Research, University of California, 1954.
20. Weertman, J., “Steady-State Creep Through Dislocation Climb,” Journal of Applied Physics,
Vol.28, pp.362-364, 1957.
21. Martin, Harold C., Graham F. Carey, Introduction to Finite Element Analysis: Theory and
Application, McGraw-Hill Book Company, New York, pp. 1-3, 1973.
22. Reddy, J. N., An Introduction to the Finite Element Method 2nd Edition. McGraw-Hill,
Boston, MA, pp. 3-15, 70-95, 1993.
23. Cook, Robert D., Concepts and Applications of Finite Element Analysis, John Wiley and
Sons, Inc., New York, pp. 1-6, 1974.
24. Nath, B., Fundamentals of Finite Elements for Engineers, The Athlone Press, London, pp.1-
6, 1974.
25. Cloyd, James S., “A Status of the United States Air Force’s More Electric Aircraft
Initiative”, Aerospace Power Systems and Technologies Proceedings of the Intersociety
Energy Conversion and Engineering Conference Vol. 1, p. 681-686, 1997.
26. Wahl, A. M., G. O. Sankey, M. J. Manjoine and E. Shoemaker, “Creep Tests of Rotating
Disks at Elevated Temperature and Comparison With Theory,” Journal of Applied
Mechanics, September 1954, pp. 225-235.
27. Wahl, A. M., “Analysis of Creep in Rotating Disks Based on the Tresca Criterion and
Associated Flow Rule,” Journal of Applied Mechanics, June 1956, pp.231-238.

140
28. Wahl, A. M., “Stress Distributions in Rotating Disks Subjected to Creep at Elevated
Temperature,” Journal of Applied Mechanics, June 1957, pp. 299-305.
29. Wahl, A. M., “Further Studies of Stress Distribution in Rotating Disks and Cylinders Under
Elevated-Temperature Creep Conditions,” Journal of Applied Mechanics, June 1958, pp.243-
250.
30. Wahl, A. M., “Effects of the Transient Period in Evaluating Rotating Disk Tests Under Creep
Conditions,” Journal of Basic Engineering, March 1963, pp.66-70.
31. Gemma, A. E., G. H. Rowe and R. J. Spahl, “Elastic and Creep Characteristics of a Class of
Shell Closures With Constant Stress Ratio”, Journal of Basic Engineering, Series D,
Transactions of the ASME, Vol. 83, 1961, p.545-550.
32. D’Isa, F. A., Analysis of Creep in Rotating Disks by The Strain Ratio Method – With Tests
on Chemically Pure Lead, University of Pittsburgh, 1960.
33. Mendelson, A., M. H. Hirschberg and S. S. Manson, “A General Approach to the Practical
Solution of Creep Problems”, Transactions of the ASME, Vol. 81, 1959, p.585-598.
34. Mendelson, A., Plasticity: Theory and Application, The Macmillian Co., New York, 1968,
p.335-345.
35. Yang, Tsung-Wen, Chung-Kuang Chen and Hwa-Kuang Shee, “Endochronic Visocelastic
Creep Analysis of 2-D Structures”, Computers and Structures Vol. 26, No. 3, pp. 425-429,
1987
36. Crossman, F.W. and M.F. Ashby, “The Nonuniform Flow of Polycrystals by Grain Boundary
Sliding Accommodated by Power Law Creep,” Acta Metallurgica, Vol. 23, No. 4, pp. 425-
440, 1975.
37. Cullity, B. D., Introduction to Magnetic Materials, Addison-Wesley Publishing Company,
Inc., Reading, MA, 1972.
38. Fingers, Richard T, Jack E.Coate and Norman E. Dowling, “Mechanical Properties of Iron-
Cobalt Alloys for Power Applications”, Aerospace Power Systems and Technology
Proceedings of the Intersociety Energy Conversion Engineering Conference Vol. 1, 1997, p.
563-568.
39. Fingers, R. T., and Kozlowski, G., “Microstructure and Magnetic Properties of Fe-Co
Alloys,” Journal of Applied Physics, Vol. 81, No. 8, p.410-411, 1997

141
40. Juvinal, Robert C., Engineering Considerations of Stress, Strain and Strength, McGraw-Hill
Book Co., pp.118-119, 128-130, 133-134
41. * Timoshenko, S., Applied Elasticity, Westing House Technical Night School Press,
Pittsburg, PA, 1925, pp. 301-311.
42. * Timoshenko, S., Strength of Materials Part II Advanced Theory and Problems, Huntington,
New York, pp. 228-235.
43. * Seel, Fred B., James O. Smith, Advanced Mechanics of Materials 2nd edition, John Wiley
and Sons, Inc., New York, pp. 304-306, 321-327.
44. * Evans, R. W., Wilshire, B., Creep of Metals and Alloys, Dostesois Printers Ltd., Bradford-
on-Avon, Wiltshire, pp. 1-27, 1985.
45. * Peterson, R. E., Stress Concentration Factors, John Wiley and Sons, Inc., New York,
pp.262-263, 291, 1974.
46. http://www.matweb.com
47. Wood, R. A., R. J. Favor, Titanium Alloys Handbook, Metals and Ceramics Information
Center, Wright-Patterson Air Force Base, Ohio, 1972.
48. * Freeman, J. W., D. N. Frey, E. E. Reynolds, and A. E. White, “Super Creep-Resistant
Alloys”, American Society for Testing Materials Symposium on Plasticity and Creep of
Metals, October 10, 1949, p. 51-68.
49. * Cross, H. C. and L. R. Jackson, “The Use of Creep Data in Design”, American Society for
Testing Materials Symposium on Plasticity and Creep of Metals, October 10, 1949, p. 37-47.
50. * Odenèo, H., “Transient Thermal Stresses in Discs With a Temperature Dependant Yield
Stress”, Acta Polytechnica Scandinavica Physics Including Nucleonics Series No. 66,
Stockholm, 1969.
51. Fingers, R. T., E-mails and discussions with

* Not referenced in the text of this report

142
6 Appendix A

6.1 INTRODUCTION
The codes genmesh.F and postp.F were created to simplify the analysis of electro-magnetic
materials for use in switched reluctance motors and generators. The code utilizes a user defined
input file which defines the material, geometric, and operating environment properties. These
codes can, in conjunction with ABAQUS, model various rotor geometries for elastic and creep
conditions. The code genmesh.F takes the user defined input file and creates an input file for
ABAQUS. In turn, postp.F translates the ABAQUS output file into files which can be used in
any spread sheet program with plotting capabilities to create time and temperature histories for
stresses and deformations. Post.F also generates files that can be used in PATRAN to generate
three dimensional deformation plots as well as stress fringe plots.

These codes along with ABAQUS can be utilized to analyze potential rotor designs and
determine designs that merit further investigation. The user should keep in mind that these codes
are not exact tools.

One final note the term “card” is used in this manual in reference to a line of input. The cards
are then grouped into categories such as “lamina variables”, “shaft variables”, etc. The term
“card” is reminiscent of the early method of data input for computers.

143
6.2 INPUT FILES
There are three basic types of files genmesh is capable of creating. The first type (type=1) will
create a model of the shaft-rotor system without the initial stress field from the interference fit.
The second type of file (type=2) will model the shaft-rotor system with the initial stress field.
This initial stress field is found by modeling a type=3 run, therefore the type=3 model must be
run prior to running a type=2 run. The input files for type=1 and type=2 runs are similar as seen
in the following sections.

144
6.2.1 Type=1 or 2 Input Files

6.2.1.1 Heading and Time Step Variables


This group of cards is used to input the job heading as well as the time step control variables.
Appendix I has an example input file which will produce the ABAQUS input file found in
appendix II. Appendix III shows an example Heading card for type 1 and 2 models.

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:3 Output file type indicator (type) I3
EQ 1 – Run shaft-rotor model without initial
stresses from interference fit
EQ 2 – Run shaft-rotor model WITH initial stress
field from interference fit (REQUIRES output from
type=3).
1:76 Job Heading (description of type of test being A76
run)
CARD 2
1:4 Number of time steps (NUMTS) I4
5:14 Total time for run in seconds (ttm) E10.0
15:24 Suggested initial time step size (sugg) E10.0
25:34 Minimum time step size (stmin) E10.0

Notes:
• In order to run type=2 you must first run type=3 to obtain the stress field formed by the
interference fit.
• Use the job-heading field to identify the type of model being analyzed.
• Each card must be preceeded by a dividing line that must have a lower case “c” in the first
column. See Appendix I.

145
6.2.1.2 Lamina Variables
This group of cards contains the information necessary to generate the nodal locations, element
connectivity, and material properties for the rotor stack comprised of a specified number of
laminas. Figure 1.1 depicts a possible configuration of a lamina stack. Appendix II shows an
example ABAQUS input file. Appendix IV has example lamina input cards for type 1 and 3
models.

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:5 Number of laminae (NUMLAM) (max = 3) I5
(figure1.1)
6:10 RADVAR I5
EQ 0 – No Radial Variation in Temperature
EQ 1 – Model Radial Variation in Temperature
11:15 CLAMP I5
EQ 0 – Inner radius free to deform
EQ 1 – Inner radius clamped (only for
multiple laminae cases NUMLAM > 1)
16:25 Inner radius (R1), inches (figure1.1) E10.0
26:35 Hub radius (R2), inches (figure1.1) E10.0
36:45 Lamina Thickness (T), inches (figure1.1) E10.0
46:48 NDEGS I3
EQ 90 – Models a quarter disk (90 degrees)
49:53 Number of elements in the radial direction I5
(NRADSEG). See figure1.2
54:58 Number of elements in the theta direction I5
(NTHSEG). See figure 1.2
59:63 Scale factor (SCALE) I5
EXAMPLE: EQ 10 ⇒ 10 units = 1 inch
* CARD 2 * Include if RADVAR EQ 1
* Repeat for each lamina (1 to NUMLAM)
1:10 (TEMPIN)Temperature at the inner radius of E10.0
the rotor (K)
11:20 (TEMPOUT)Temperature at the outer radius of E10.0
the rotor (either hub if no poles are
modeled or pole-tips) (K)

146
COLUMN(S) DESCRIPTION FORMAT
CARD 3
1:2 Lamina material type indicator (ELCRPL) I2
EQ 0 – Isotropic Elastic Material
EQ 1 – Transient Thermal Creep Material
3:80 Lamina material title (MTLTITL) A78
** CARD 4 ** Include if ELCRPL EQ 0 and RADVAR EQ 0
1:10 Density (RHOL) , (lb s^2/in^4) E10.0
11:20 Young’s modulus (YMODL), (psi) E10.0
21:30 Poisson’s ratio (NUL) E10.0
*** CARD 5 *** Include if ELCRPL EQ 0 and RADVAR EQ 1
1:10 (MM) Slope of Elastic Modulus Curve. See E10.0
Figure 1.3 (psi)
11:20 (B) Vertical (Y) intercept of Elastic E10.0
Modulus Curve. See Figure 1.3
21:30 (RHOL(1)) Density (lb s^2/in^4) E10.0
31:40 (NUL(1)) Poisson’s ratio. E10.0
**** CARD 6 **** Include if ELCRPL EQ 1 and RADVAR EQ 0
**** Repeat combination of CARD 6 & 7 for
each lamina
1:10 Density (RHOL), (lb s^2/in^4) E10.0
11:20 Poisson’s ratio (NUL) E10.0
**** CARD 7
1:10 Temperature (T1), (K) E10.0
11:20 Young’s Modulus (Ee), (psi) E10.0
21:30 Stress coefficient (A), (1/(s (psi)^n)) E10.0
31:40 Stress exponent (N) E10.0
***** CARD 8 ***** Include if ELCRPL EQ 1 and RADVAR EQ 1
1:10 (MM) Slope of Elastic Modulus Curve. See E10.0
Figure 1.3 (psi/F)
11:20 (B) Vertical (Y) intercept of Elastic E10.0
Modulus Curve. See Figure 1.3 (psi)
21:30 (RHOL(1)) Density (lb s^2/in^4) E10.0
31:40 (NUL(1)) Poisson’s ratio. E10.0
***** CARD 9
1:10 (AT) Stress Exponent (K/s (psi)^Nn) E10.0
11:20 (Q) Activation Energy (cal/mol) E10.0
21:30 (R) Universal Gas Constant (cal/(mol K)) E10.0
31:40 (Nn) Stress Exponent E10.0

147
COLUMN(S) DESCRIPTION FORMAT
****** CARD 10 ****** Include if numlam > 1
1:3 Friction indicator E10.0
EQ 1 – Frictionless sliding
EQ 2 – Sliding with friction
11:20 Coefficient of static friction (MUS) E10.0

Notes:
• The maximum number of lamina is currently three, if there is a need for more laminae the
code will need to be altered accordingly.
• The code has the capability of modeling a rotor with four poles (each pole separated by 90
degrees), therefore the variable ndegs on card one should be equal to 90. If more than 4
poles are to be modeled the code will have to be modified.
• The lamina material title on card two is just a means to make the input file easier to read, this
variable is not used in the generation of the ABAQUS input file.
• The variable elcrpl indicates the type of material model being used for the laminae stack. If
elcrpl equals zero then an elastic model is assumed and card 4 or 5 should be entered. This
card should be repeated for each lamina. If elcrpl equals one then a transient thermal creep
model is assumed. The combination of card 6 and card 7 or card 8 and card 9 should be
entered. Cards 6 and 7 should each be repeated for every lamina (e.g. For lamina one input
card 6 and then seven, for lamina two input card 6 and then 7, etc…).
• The friction card should only be entered if at least two laminae are being modeled, otherwise
do not include this card.
• The code has been set up to handle a stack of lamina each with a different radial variation in
temperature. Thereby, allowing the modeling of T(r,z). This aspect was not fully tested
and most likely needs further work. The code is capable of handling a stack of different
constant temperature lamina.

148
Figure 1.1 – Force and Boundary Conditions acting on the Shaft-Rotor System

Figure 1.2 – Mesh Generation Variables

149
Figure 1.3 – Variable Temperature Profile

150
6.2.1.3 Pole Variables
This card contains the information necessary to generate the pole nodes and element
connectivity. Appendix I contains an example input file. Appendix V contains example pole
cards for type 1, 2 and 3 models.

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:3 Pole indicator (py) I3
EQ 0 – poles are not modeled
EQ 1 – include poles in model (currently only
4 poles can be modeled, each separated by 90
degrees)
4:13 Pole-tip radius (r3), see figure 1.2 E10.0
14:23 Half pole width ( ), see figure 1.2 E10.0
24:28 Number of elements in the radial direction I5
(nthpl), see figure 1.2

151
6.2.1.4 Shaft Variables
This group of cards contains all the information necessary to generate the nodal locations,
element connectivity, and material properties for the shaft, upon which the rotor stack is
mounted. Figure 1.3 depicts a possible configuration of a shaft. Appendix II shows an example
of an ABAQUS input file. Appendix VI has example shaft input cards for type 1 and 2 models.

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:3 Shaft indicator (TY) I3
EQ 0 – Shaft is not modeled
EQ 1 – include shaft in model (do NOT use with
multiple laminae option)
4:13 Wall thickness of the shaft (TBT), inches. E10.0
Note that the outer radius is the same as the
inner radius of the lamina stack (figure1.1)
14:18 Number of elements in the radial direction for I5
the shaft (NTBSEG). Note that the number of
elements in the theta direction is the same as
used in the lamina stack(figure1.2)
* CARD 2 * If TY EQ 1
1:2 Shaft material type indicator (ELCRPT) I2
EQ 0 – Isotropic Elastic Material
EQ 1 – Transient Thermal Creep Material
3:71 Shaft material title (MTLTITT) A69
** CARD 3 ** Include if ELCRPT EQ 0 and TY EQ 1
1:10 Density (RHOT) , (lb s^2/in^4) E10.0
11:20 Material reference temperature (MTRFTT), (R) E10.0
21:30 Young’s modulus (YMODT), (psi) E10.0
31:40 Poisson’s ratio (NUT) E10.0
*** CARD 4 *** Include if ELCPT EQ 1 and TY EQ 1
1:10 Density (RHOT), (lb s^2/in^4) E10.0
11:20 Stress exponent (N) E10.0
*** CARD 5
1:10 Temperature (T1), (K) E10.0
11:20 Young’s Modulus (Ee), (psi) E10.0
21:30 Stress coefficient (A), (1/(s (psi)^n)) E10.0

152
Notes:
• If ty equals zero then cards 2 and beyond are not necessary and should not be added to the
input file.

6.2.1.5 Load Variables


This card contains the information necessary to generate the load curve, concentrated nodal load
and pressure input for ABAQUS which will model all of the forces acting on the rotor-shaft
model (including force due to rotation, friction, and the axial clamping force). See appendix II
for an example ABAQUS input file. Figure 1.4 demonstrates the forces acting on the rotor-shaft
model. Appendix VII has an example load input card.

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:3 Angular velocity indicator (WY) I3
EQ 0 – No rotation
EQ 1 – Rotation is modeled
4:6 Axial clamping force indicator (AXY) I3
EQ 0 – No clamping force
EQ 1 – Clamping force is modeled
7:16 Angular velocity (ANGVEL), rpm E10.0
See figure 1.4
17:26 Axial Clamping Force (SIGMAZ), psi E10.0
See figure 1.4

153
6.2.2 Type = 3 Input File

6.2.2.1 Heading and Interference Fit Variables


This group of cards is used to input the job heading as well as the interference fit variables. The
equivalent pressure required to generate the interference fit stress field is computed by the code.
Appendix VIII has an example type three input file.

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:3 Output file type indicator (type) I3
EQ 3 – Run model to generate interference fit
stress field
1:76 Job Heading (description of type of test being A76
run)
CARD 2
1:10 Interference Fit (intfit), inches E10.0

Notes:
• The equivalent interference fit pressure is calculated by the method outlined by Juvinal [40].
• Use the job-heading field to identify the type of model being analyzed.
• Each card must be preceeded by a dividing line that must have a lower case “c” in the first
column. See Appendix I.

154
6.2.2.2 Lamina Variables
These cards contain the information necessary to generate the nodal locations, element
connectivity, and material properties for the rotor stack comprised of a specified number of
laminas. Figure 1.1 depicts a possible configuration of a laminae stack. Appendix VIII has an
example type three input file.

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:10 Inner radius (R1), inches (figure1.1) E10.0
11:20 Hub radius (R2), inches (figure1.1) E10.0
21:30 Lamina Thickness (T), inches (figure1.1) E10.0
31:33 NDEGS I3
EQ 90 – Models a quarter disk (90 degrees)
EQ 360 – Models a whole disk (360 degrees)
(not fully coded)
34:38 Number of elements in the radial direction I5
(NRADSEG). See figure1.2
39:43 Number of elements in the theta direction I5
(NTHSEG). See figure 1.2
44:48 Scale factor (SCALE) I5
EXAMPLE: EQ 10 ⇒ 10 units = 1 inch
CARD 2
1:10 Density (RHOL) , (lb s^2/in^4) E10.0
11:20 Young’s modulus (YMODL), (psi) E10.0
21:30 Poisson’s ratio (NUL) E10.0

Notes:
• For a type=3 run only one lamina is modeled therefore there is no need for a number of
laminas variable.
• The code has the capability of modeling a rotor with four poles (each pole separated by 90
degrees), therefore the variable ndegs on card one must be equal to 90. If more than 4 poles
are to be modeled the code will have to be modified.

155
6.2.2.3 Pole Variables
This card contains the information necessary to generate the pole nodes and element
connectivity. Appendix VIII has an example type three input file. Appendix V contains example
pole cards for type 1, 2 or 3 models..

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:3 Pole indicator (py) I3
EQ 0 – poles are not modeled
EQ 1 – include poles in model (currently only
4 poles can be modeled, each separated by 90
degrees)
4:13 Pole-tip radius (r3), see figure 1.2 E10.0
14:23 Half pole width ( ), see figure 1.2 E10.0
24:28 Number of elements in the radial direction I5
(nthpl), see figure 1.2

156
6.2.2.4 Shaft Variables

This card contains all the information necessary to generate the nodal locations, element
connectivity, and material properties for the shaft, upon which the rotor stack is mounted. Figure
1.3 depicts a possible configuration of a shaft. Appendix VIII has an example type three input
file. Appendix VI has an example shaft card for type = 1, 2 and 3 models.

COLUMN(S) DESCRIPTION FORMAT


CARD 1
1:3 Shaft indicator (TY) I3
EQ 0 – Shaft is not modeled
EQ 1 – include shaft in model
4:13 Wall thickness of the shaft (TBT), inches. E10.0
Note that the outer radius is the same as the
inner radius of the lamina stack (figure1.1)
14:18 Number of elements in the radial direction for I5
the shaft (NTBSEG). Note that the number of
elements in the theta direction is the same as
used in the lamina stack (figure1.2)
* CARD 2 * If TY EQ 1
1:10 Density (RHOT) , (lb s^2/in^4) E10.0
11:20 Material reference temperature (MTRFTT), (R) E10.0
21:30 Young’s modulus (YMODT), (psi) E10.0
31:40 Poisson’s ratio (NUT) E10.0

Notes:
• If ty equals zero then card 2 is not necessary and should not be added to the input file.

157
6.3 Running genmesh.F
Genmesh.F is a Unix based Fortran 77 code which utilizes the user generated input file (in the
format specified above) in order to generate an input file for ABAQUS. An example ABAQUS
input file is located in appendix I (Please note that the exact format of the ABAQUS input file
will vary depending on the selections made in the user generated input file).

Once the user generated input file has been created, genmesh.F must be run. Genmsh.F is the
source code and, therefore, needs to be complied prior to execution. A Fortran 77 Unix based
compiler is required. To compile simply type in the directory which genmesh.F is found: f77
genmesh.F. This will compile the code and generate an executable file named a.out. I usually
rename the file to genmesh.exe. Please note that the code only needs to be compiled once after
that it should only be done if any changes have been made.

Once genmesh.F is complied simply run the code by typing in the name of the executable created
(genmesh.exe). The user generated input file must be named input (note that no extension is
used) and it must be located in the same directory as genmesh.exe. Genmesh.exe will then
generate the ABAQUS input file entitled model.inp for type equal to one or two runs. The files
shaft.inp and rotor.inp will be created for type equal to three runs. These files can be renamed
but they must have the extension “inp”(Please note that if one renames the files prior to post
processing that the post processor will not function since by default it looks for these file names).
Now ABAQUS needs to be run. It is highly recommended that each run be preformed in a
separate directory, this will help keep track of all the files generated for each model. If this
method is used then the directory name can be used to differentiate the various models, rather
than renaming the files upon the completion of the analysis.

Note, if a type two model is being created then the type three results must have already been
obtained. The type two model will utilize the files shaft.dat and rotor.dat to generate the stress
initial conditions representing the interference fit.

158
6.4 Running ABAQUS
ABAQUS must be installed on the workstation. Then enter the directory containing the
ABAQUS input file, model.inp. To run the input file simply abaqus job=model, if you have
changed the name of the input file just replace “model” with the name of the input file minus the
inp extension. This will generate many files among which are: model.dat, model.log, model.msg
and model.fil (if a type three model is being analyzed then the same files would be created but
with shaft and rotor instead of model). The log file contains a log of the input and calculation
process, any errors will be indicated here. The msg file contains detailed convergence
information, any error during the calculation process will be detailed here. The dat file contains
the user-requested output. The fil file contains the user requested output in a format which is
transferable to other platforms (programs and operating systems).

159
6.5 Running postp.F and creating results files
Postp.F is a post processor that will compile the data generated by ABAQUS and create a few
files that can be used to analyze the results. These files were created with the use of Microsoft
Excel and PATRAN in mind (note that the files generated for Excel can also be used with any
spreadsheet program). The first step is to copy the ABAQUS output file (the .dat files) into the
same directory which houses postp.F. Postp.F once again denotes the source code which can be
complied in the same manner as indicated for genmesh.F. Once postp.F is complied, run the
executable (postp.exe).

The code will prompt the user to select the type run being analyzed. To analyze type 1 or two
runs select option 1. To analyze type 3 runs select option 2.

Next the user is prompted to choose the type of files to generate. These types of files are defined
as follows:

model_dis.exc – Spreadsheet radial displacement file. Used to generate radial displacement


profiles. To generate select option 1 or 3
model.dv(1,2,3) – Spreadsheet radial displacement vs. time file. The user can indicate up to three
different radial positions to generate files for radial displacement vs. time plots. To generate
select option 1 or 3 and indicate yes when prompted to generate radial displacement vs. time
files.
model_sts.exc – Spreadsheet stress file. Used to generate stress, strain and strain rate profiles.
To generate select option 2 or 3.

Note, when postp.exe is run it will prompt the user to indicate the run type being analyzed. If
type one or two results are being post processed then the code will use the ABAQUS result file
model.dat, if a type three run is being analyzed the result files shaft.dat and rotor.dat will be
utilized and the files generated will begin with shaft and rotor rather than model.

160
Deformation and fringe plots can also be created using the ABAQUS model.fil file. This file is
designed to be portable to other programs and operating systems. I used PATRAN version 9.0 to
generate the deformations and fringe plots. This is accomplished by starting PATRAN and
creating a new file. You will be prompted to indicate the analysis type when selecting the
properties for the new file, make sure to select ABAQUS. Once the new file is open, select
analysis from the tool bar. Then select read result, and both to read in both the model data and
results. Finally select the results file and click on apply. This will read in the model.fil file.
Once this is complete the plots can be created using the results option on the tool bar.

161
6.6 Appendix

6.6.1 Appendix I – Example Type = 1 and 2 input file (input)


c---5---10---15---20---25---30---35---40---45---50---55---60---65---70---75---80
1 One lamina,T=780K,d=0.0015,t=5000hrs,w=25000rpms,Ati=AFeCo
5000 1.800E+07 1.000E-03 1.000E-08
c-------------------Lamina Variables------------------------------------------
1 0 0 0.500E+00 1.730E+00 6.000E-03 90 6 12 1
1 Creep Material (FeCo)
7.588E-04 3.000E-01
7.800E+02 1.991E+07 3.535E-27 3.830E+00
c-------------------Pole Variables----------------------------------------------
1 2.150E+00 5.500E-01 10
c-------------------Shaft Variables---------------------------------------------
1 1.250E-01 10
1 Creep Material (Ti)
4.206E-04 3.400E-01
7.800E+02 1.680E+07 3.535E-27 3.830E+00
c-------------------Load Variables----------------------------------------------
1 0 2.500E+04 0.000E+00

Notes:
• Each group of cards is separated by a dividing line, which contains a “c” in the first column.
• This input file will produce the ABAQUS input file found in appendix II.

162
6.6.2 Appendix II – Example ABAQUS input file (model.inp)
*HEADING
1 lamina,T=780K,d=0.0015,t=5000hrs,w=25000rpms,Ati=AFeCo
*PREPRINT,ECHO=YES,HISTORY=NO,MODEL=NO
*NODE
1, 3.7500000E-01, 0.0000000E+00, 0.0000000E+00
2, 3.7282140E-01, 4.0363388E-02, 0.0000000E+00
3, 3.6631092E-01, 8.0257786E-02, 0.0000000E+00
4, 3.5554419E-01, 1.1921965E-01, 0.0000000E+00
5, 3.3305693E-01, 1.7233131E-01, 0.0000000E+00
6, 3.0269111E-01, 2.2136642E-01, 0.0000000E+00
7, 2.6516504E-01, 2.6516504E-01, 0.0000000E+00
8, 2.2136642E-01, 3.0269111E-01, 0.0000000E+00
9, 1.7233131E-01, 3.3305693E-01, 0.0000000E+00
10, 1.1921965E-01, 3.5554419E-01, 0.0000000E+00
11, 8.0257786E-02, 3.6631092E-01, 0.0000000E+00
12, 4.0363388E-02, 3.7282140E-01, 0.0000000E+00
13, -3.8862599E-14, 3.7500000E-01, 0.0000000E+00
.
.
.
628, 0.0000000E+00, 2.1500000E+00, 6.0000000E-03
*ELEMENT,TYPE=C3D8R,ELSET=TBIN1
1, 1, 14, 15, 2, 144, 157, 158, 145
2, 2, 15, 16, 3, 145, 158, 159, 146
3, 3, 16, 17, 4, 146, 159, 160, 147
4, 4, 17, 18, 5, 147, 160, 161, 148
5, 5, 18, 19, 6, 148, 161, 162, 149
.
.
.
*ELEMENT,TYPE=C3D8R,ELSET=TOUT1
109, 118, 131, 132, 119, 261, 274, 275, 262
110, 119, 132, 133, 120, 262, 275, 276, 263
111, 120, 133, 134, 121, 263, 276, 277, 264
112, 121, 134, 135, 122, 264, 277, 278, 265
113, 122, 135, 136, 123, 265, 278, 279, 266
114, 123, 136, 137, 124, 266, 279, 280, 267
115, 124, 137, 138, 125, 267, 280, 281, 268
116, 125, 138, 139, 126, 268, 281, 282, 269

163
117, 126, 139, 140, 127, 269, 282, 283, 270
118, 127, 140, 141, 128, 270, 283, 284, 271
119, 128, 141, 142, 129, 271, 284, 285, 272
120, 129, 142, 143, 130, 272, 285, 286, 273
*ELSET,ELSET=TB1
TBIN1,TOUT1
*ELSET,ELSET=TUBE
TBIN1,TOUT1
*ELEMENT,TYPE=C3D8R,ELSET=LAM1
121, 287, 300, 301, 288, 458, 471, 472, 459
122, 288, 301, 302, 289, 459, 472, 473, 460
123, 289, 302, 303, 290, 460, 473, 474, 461
124, 290, 303, 304, 291, 461, 474, 475, 462
125, 291, 304, 305, 292, 462, 475, 476, 463
.
.
.

252, 448, 456, 457, 449, 619, 627, 628, 620


*ELSET,ELSET=LM1
LAM1,LOT1
*SOLID SECTION,MATERIAL=TBMAT,ELSET=TUBE
*SOLID SECTION,MATERIAL=FECO1,ELSET=LM1
*SURFACE DEFINITION,NAME=MAST1
TOUT1,S4
*SURFACE DEFINITION,NAME=SLAV1
LAM1,S6
*SURFACE INTERACTION,NAME=TDD
*CONTACT PAIR,INTERACTION=TDD,TIED,ADJUST= 1.250E-05
SLAV1,MAST1
*MATERIAL,NAME=TBMAT
*ELASTIC,TYPE=ISOTROPIC
1.680E+07, 3.400E-01
*CREEP,LAWS=TIME
3.535E-27, 3.830E+00,0.
*MATERIAL,NAME=FECO1
*ELASTIC,TYPE=ISOTROPIC
1.991E+07, 3.000E-01
*CREEP,LAWS=TIME
3.535E-27, 3.830E+00,0.
*BOUNDARY
1,2,3

164
2,3
3,3
4,3
5,3
6,3
.
.
.
628,1
*STEP,INC=5000
*VISCO,CETOL=5.E-5
1.000E-03, 1.800E+07, 1.000E-08
*CLOAD
1,1, 4.17425400E-04
1,2, 1.50057000E-05
2,1, 8.30000600E-04
2,2, 8.98597000E-05
3,1, 8.15506500E-04
3,2, 1.78675400E-04
.
.
.
628,1, 1.89711320E-03
628,2, 6.51603706E-02
*PRINT,FREQ=10
*EL FILE,FREQ=10
S,
COORD,
E,
*NODE FILE,FREQ=10
U,COORD
*NODE PRINT,FREQ=10
U,COORD
*EL PRINT,FREQ=10
S,
COORD,
E,
*END STEP

165
6.6.3 Appendix III – Example Heading Cards
c---5---10---15---20---25---30---35---40---45---50---55---60---65---70---75---80
1 One lamina,T=780K,d=0.0015,t=5000hrs,w=25000rpms,Ati=AFeCo
5000 1.800E+07 1.000E-03 1.000E-08

166
6.6.4 Appendix IV – Example Lamina Cards
c-------------------Lamina Variables------------------------------------------
1 0 0 0.500E+00 1.730E+00 6.000E-03 90 6 12 1
1 Creep Material (FeCo)
7.588E-04 3.000E-01
7.800E+02 1.991E+07 3.535E-27 3.830E+00

Notes:
• This group of lamina cards indicates that one lamina is modeled.
• The radial variation option (RADVAR) is zero indicating no radial variation in temperature
• The option CLAMP is zero indicating that the inner radius is free to deform
• The material model used is the Creep model.
• Since only one lamina is modeled, there is only one set of material variables (the last two
card).

c-------------------Lamina Variables------------------------------------------
2 1 1 0.500E+00 1.730E+00 6.000E-03 90 6 12 1
5.880E+02 8.100E+02
5.880E+02 8.100E+02
1 Creep Material (FeCo)
-.104E+05 2.973E+07 7.588E-04 3.000E-01
4.940E+07 9.885E+04 1.987E+00 3.830E+00
2 0.200E+00

Notes:
• This group of cards indicates two laminas are modeled.
• RADVAR EQ 1 indicating that there is a radial variation in temperature
• CLAMP EQ 1 indicating that the inner radius is clamped
• Notice that the radial temperature profiles are the same for each lamina.
• Since there are more than one lamina, the friction card must be included. In this case the
card indicates frictionless sliding (the last card).

167
6.6.5 Appendix V – Example Pole cards
c-------------------Pole Variables---------------------------------------------
1 2.500E+00 5.000E-01 4

6.6.6 Appendix VI – Example Shaft Cards


c-------------------Shaft Variables----------------------------------------------
0 0.000E+00 0

Notes:
• Since ty = 0 the rest of the cards used to define the material properties of the shaft are not
necessary.

c-------------------Shaft Variables----------------------------------------------
1 1.250E-01 3
1 Creep Material
7.588E-04 3.000E-01
8.109E+02 1.833E+07 3.841E-26 3.830E+00

Notes:
• If the shaft is modeled, the material property cards are then added.

6.6.7 Appendix VII – Example Load card


c-------------------Load Variables----------------------------------------------
1 1 5.500E+04 5.000E+02

168
6.6.8 Appendix VIII – Example Type = 3 Input File (input)
c---5---10---15---20---25---30---35---40---45---50---55---60---65---70---75---80
3 Interference Fit Test
1.500E-03
c-------------------Lamina Variables------------------------------------------
0.500E+00 1.730E+00 6.000E-03 90 20 60 1
7.588E-04 2.702E+07 3.000E-01
c-------------------Pole Variables----------------------------------------------
1 2.150E+00 5.500E-01 4
c-------------------Tube Variables----------------------------------------------
1 1.250E-01 4
4.206E-04 1.680E+07 3.400E-01

169
7 Appendix B – Creep Type A Results

Von Mises Stain Fringe Plot at 5000hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-1 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=35,000rpm,
Type A, ATi=AfeCo

170
Von Mises Stain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-2 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=40000rpm,
Type A, ATi=AfeCo

171
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-3 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=45,000rpm,
Type A, ATi=AFeCo

172
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-4 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=45,000rpm,
Type A, ATi=AFeCo

173
Von Mises Strain Fringe Plot at 5000hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-5 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=50,000rpm,
Type A, ATi=AFeCo

174
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-6 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=55,000rpm,
Type A, ATi=AFeCo

175
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-7 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=30000rpm,
Type A, ATi=AFeCo

176
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-8 - Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=35,000rpm,
Type A, ATi=AFeCo

177
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-9 - Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=35,000rpm,
Type A, ATi=AFeCo

178
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-10 - Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=40000rpm,
Type A, ATi=AFeCo

179
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-11 - Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=25,000rpm,
Type A, ATi=AFeCo

180
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-12 - Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=27000rpm,
Type A, ATi=AFeCo

181
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-13 - Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=30000rpm,
Type A, ATi=AFeCo

182
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-14 - Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=25,000rpm,
Type A, ATi=AFeCo

183
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-15 - Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=30000rpm,
Type A, ATi=AFeCo

184
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-16 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=588K,
Tout=810K, ω=45,000rpm, Type A

185
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-17 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=588K,
Tout=810K, ω=47000rpm, Type A

186
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-18 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=588K,
Tout=810K, ω=48500rpm, Type A

187
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-19 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=727K,
Tout=810K, ω=30000rpm, Type A

188
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-20 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=727K,
Tout=810K, ω=35,000rpm, Type A

189
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-21 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=727K,
Tout=810K, ω=40000rpm, Type A

190
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-22 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=810K,
Tout=588K, ω=25,000rpm, Type A

191
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-23 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=810K,
Tout=588K, ω=27000rpm, Type A

192
\Von Mises Stress Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure B-24 - Results for Radially Varying Temperature Shaft-Rotor System for d=0.0015’’, Tin=810K,
Tout=588K, ω=30000rpm, Type A

193
8 Appendix C – Creep Type AB1 Results

Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-1 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=35,000rpm,
Type AB1, ATi=AFeCo

194
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-2 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=40000rpm,
Type AB1, ATi=AFeCo

195
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-3 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=45,000rpm,
Type AB1, ATi=AFeCo

196
Von Mises Strain Fringe Plot at 75 hrs

Von Mises Stress Fringe Plot at 75 hrs

Figure C-4 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=727K, ω=50,000rpm,
Type AB1, ATi=AFeCo

197
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-5 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=45,000rpm,
Type AB1, ATi=AFeCo

198
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-6 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=50,000rpm,
Type AB1, ATi=AFeCo

199
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-7 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=55,000rpm,
Type AB1, ATi=AFeCo

200
Von Mises Strain Fringe Plot at 1557 hrs

Von Mises Stress Fringe Plot at 1557 hrs

Figure C-8 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=727K, ω=60000rpm,
Type AB1, ATi=AFeCo

201
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-9 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=30,000rpm,
Type AB1, ATi=AFeCo

202
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-10 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=35,000rpm,
Type AB1, ATi=AFeCo

203
Von Mises Strain Fringe Plot at 1910 hrs

Von Mises Stress Fringe Plot at 1910 hrs

Figure C-11 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=760K, ω=40000rpm,
Type AB1, ATi=AFeCo

204
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-12 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=35,000rpm,
Type AB1, ATi=AFeCo

205
Von Mises Strain fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-13 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=40,000rpm,
Type AB1, ATi=AFeCo

206
Von Mises Strain Fringe Plot at 530 hrs

Von Mises Stress Fringe Plot at 530 hrs

Figure C-14 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=760K, ω=45,000rpm,
Type AB1, ATi=AFeCo

207
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-15 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=25,000rpm,
Type AB1, ATi=AFeCo

208
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-16 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=30,000rpm,
Type AB1, ATi=AFeCo

209
Von Mises Strain Fringe Plot at 932 hrs

Von Mises Stress Fringe Plot at 932 hrs

Figure C-17 – Results for Constant Temperature Shaft-Rotor System for d=0.0015’’, T=780K, ω=35,000rpm,
Type AB1, ATi=AFeCo

210
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-18 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=25,000rpm,
Type AB1, ATi=AFeCo

211
Von Mises Strain Fringe Plot at 5000 hrs

Von Mises Stress Fringe Plot at 5000 hrs

Figure C-19 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=30000rpm, Type AB1,
ATi=AFeCo

212
Von Mises Strain Fringe Plot at 588 hrs

Von Mises Stress Fringe Plot at 588 hrs

Figure C-20 – Results for Constant Temperature Shaft-Rotor System for d=0.003’’, T=780K, ω=35,000rpm, Type AB1,
ATi=AfeCo

213
9 VITA

Donald Andrew Phillips was born in Falls Church Virginia on April 13th, 1974 to MacDonald and Inés
Phillips. He grew up in Northern Virginia and graduated from Woodbridge Senior High School in 1992. He
received his Bachelor of Science degree in Aerospace Engineering from Virginia Polytechnic Institute and
State University in 1997.

214

Potrebbero piacerti anche